+ All Categories
Home > Documents > Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if...

Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if...

Date post: 21-Jun-2018
Category:
Upload: buicong
View: 221 times
Download: 0 times
Share this document with a friend
43
ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS FOLKMAR BORNEMANN * Abstract. Some significant quantities in mathematics and physics are most naturally expressed as the Fredholm determinant of an integral operator, most notably many of the distribution functions in random matrix theory. Though their numerical values are of interest, there is no systematic numerical treatment of Fredholm determinants to be found in the literature. Instead, the few numerical evaluations that are available rely on eigenfunction expansions of the operator, if expressible in terms of special functions, or on alternative, numerically more straightforwardly accessible analytic expressions, e.g., in terms of Painlevé transcendents, that have masterfully been derived in some cases. In this paper we close the gap in the literature by studying projection methods and, above all, a simple, easily implementable, general method for the numerical evaluation of Fredholm determinants that is derived from the classical Nyström method for the solution of Fredholm equations of the second kind. Using Gauss–Legendre or Clenshaw– Curtis as the underlying quadrature rule, we prove that the approximation error essentially behaves like the quadrature error for the sections of the kernel. In particular, we get exponential convergence for analytic kernels, which are typical in random matrix theory. The application of the method to the distribution functions of the Gaussian unitary ensemble (GUE), in the bulk and the edge scaling limit, is discussed in detail. After extending the method to systems of integral operators, we evaluate the two- point correlation functions of the more recently studied Airy and Airy 1 processes. Key words. Fredholm determinant, Nyström’s method, projection method, trace class operators, ran- dom matrix theory, Tracy–Widom distribution, Airy and Airy 1 processes AMS subject classifications. 65R20, 65F40, 47G10, 15A52 1. Introduction. Fredholm’s (1903) landmark paper 1 on linear integral equa- tions is generally considered to be the forefather of those mathematical concepts that finally led to modern functional analysis and operator theory—see the historical ac- counts in Kline (1972, pp. 10581075) and Dieudonné (1981, Chap. V). Fredholm was interested in the solvability of what is now called a Fredholm equation of the second kind, u( x)+ z Z b a K( x, y)u(y) dy = f ( x) ( x (a, b)), (1.1) and explicit formulas for the solution thereof, for a right hand side f and a kernel K, both assumed to be continuous functions. He introduced his now famous determi- nant d(z)= k=0 z n n! Z b a ··· Z b a det ( K(t p , t q ) ) n p,q=1 dt 1 ··· dt n , (1.2) which is an entire function of z C, and succeeded in showing that the integral equation is uniquely solvable if and only if d(z) 6 = 0. Realizing the tremendous potential of Fredholm’s theory, Hilbert started work- ing on integral equations in a flurry and, in a series of six papers from 1904 to 1910, 2 * Zentrum Mathematik, Technische Universität München, Boltzmannstr. 3, 85747 Garching, Germany ( ). Manuscript as of June 24, 2008. 1 Simon (2005, p. VII) writes: “This deep paper is extremely readable and I recommend it to those wishing a pleasurable afternoon.” An English translation of the paper can be found in Birkhoff (1973, pp. 449465). 2 Later reproduced as one of the first books on linear integral equations (Hilbert 1912). 1 arXiv:0804.2543v2 [math.NA] 24 Jun 2008
Transcript
Page 1: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS

FOLKMAR BORNEMANN∗

Abstract. Some significant quantities in mathematics and physics are most naturally expressed as theFredholm determinant of an integral operator, most notably many of the distribution functions in randommatrix theory. Though their numerical values are of interest, there is no systematic numerical treatmentof Fredholm determinants to be found in the literature. Instead, the few numerical evaluations that areavailable rely on eigenfunction expansions of the operator, if expressible in terms of special functions,or on alternative, numerically more straightforwardly accessible analytic expressions, e.g., in terms ofPainlevé transcendents, that have masterfully been derived in some cases. In this paper we close the gapin the literature by studying projection methods and, above all, a simple, easily implementable, generalmethod for the numerical evaluation of Fredholm determinants that is derived from the classical Nyströmmethod for the solution of Fredholm equations of the second kind. Using Gauss–Legendre or Clenshaw–Curtis as the underlying quadrature rule, we prove that the approximation error essentially behaveslike the quadrature error for the sections of the kernel. In particular, we get exponential convergencefor analytic kernels, which are typical in random matrix theory. The application of the method to thedistribution functions of the Gaussian unitary ensemble (GUE), in the bulk and the edge scaling limit,is discussed in detail. After extending the method to systems of integral operators, we evaluate the two-point correlation functions of the more recently studied Airy and Airy1 processes.

Key words. Fredholm determinant, Nyström’s method, projection method, trace class operators, ran-dom matrix theory, Tracy–Widom distribution, Airy and Airy1 processes

AMS subject classifications. 65R20, 65F40, 47G10, 15A52

1. Introduction. Fredholm’s (1903) landmark paper1 on linear integral equa-tions is generally considered to be the forefather of those mathematical concepts thatfinally led to modern functional analysis and operator theory—see the historical ac-counts in Kline (1972, pp. 1058–1075) and Dieudonné (1981, Chap. V). Fredholm wasinterested in the solvability of what is now called a Fredholm equation of the secondkind,

u(x) + z∫ b

aK(x, y)u(y) dy = f (x) (x ∈ (a, b)), (1.1)

and explicit formulas for the solution thereof, for a right hand side f and a kernel K,both assumed to be continuous functions. He introduced his now famous determi-nant

d(z) =∞

∑k=0

zn

n!

∫ b

a· · ·

∫ b

adet

(K(tp, tq)

)np,q=1 dt1 · · · dtn, (1.2)

which is an entire function of z ∈ C, and succeeded in showing that the integralequation is uniquely solvable if and only if d(z) 6= 0.

Realizing the tremendous potential of Fredholm’s theory, Hilbert started work-ing on integral equations in a flurry and, in a series of six papers from 1904 to 1910,2

∗Zentrum Mathematik, Technische Universität München, Boltzmannstr. 3, 85747 Garching, Germany([email protected]). Manuscript as of June 24, 2008.

1Simon (2005, p. VII) writes: “This deep paper is extremely readable and I recommend it to thosewishing a pleasurable afternoon.” An English translation of the paper can be found in Birkhoff (1973,pp. 449–465).

2Later reproduced as one of the first books on linear integral equations (Hilbert 1912).

1

arX

iv:0

804.

2543

v2 [

mat

h.N

A]

24

Jun

2008

Page 2: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

2 F. BORNEMANN

transformed the determinantal framing to the beginnings of what later, in the handsof Schmidt, Carleman, Riesz, and others, would become the theory of compact op-erators in Hilbert spaces. Consequently, over the years Fredholm determinants havefaded from the core of general accounts on integral equations to the historical re-marks section—if they are mentioned at all.3

So, given this state of affairs, why then study the numerical evaluation of theFredholm determinant d(z)? The reason is, simply enough, that the Fredholm deter-minant and the more general notions, generalizing (1.1) and (1.2) to

u + zAu = f , d(z) = det(I + zA),

for certain classes of compact operators A on Hilbert spaces, have always remainedimportant tools in operator theory and mathematical physics (Gohberg, Goldbergand Krupnik 2000, Simon 2005). In turn, they have found many significant appli-cations: e.g., in atomic collision theory (Jost and Pais 1951, Moiseiwitsch 1977), in-verse scattering (Dyson 1976), in Floquet theory of periodic differential equations(Eastham 1973), in the infinite-dimensional method of stationary phase and Feyn-man path integrals (Albeverio and Høegh-Krohn 1977, Rezende 1994), as the two-point correlation function of the two-dimensional Ising model (Wilkinson 1978), inrenormalization in quantum field theory (Simon 2005), as distribution functions inrandom matrix theory (Mehta 2004, Deift 1999, Katz and Sarnak 1999) and com-binatorial growth processes (Johansson 2000, Prähofer and Spohn 2002, Sasamoto2005, Borodin, Ferrari, Prähofer and Sasamoto 2007). As Lax (2002, p. 260) puts itmost aptly upon including Fredholm’s theory as a chapter of its own in his recenttextbook on functional analysis: “Since this determinant appears in some moderntheories, it is time to resurrect it.”

In view of this renewed interest in operator determinants, what numerical meth-ods are available for their evaluation? Interestingly, this question has apparentlynever—at least to our knowledge—been systematically addressed in the numericalanalysis literature.4 Even experts in the applications of Fredholm determinants com-monly seem to have been thinking (Spohn 2008) that an evaluation is only possible

3For example, Hilbert (1912), Webster (1927), and Whittaker and Watson (1927) start with the Fred-holm determinant right from the beginning; yet already Courant and Hilbert (1953, pp.142–147), thetranslation of the German edition from 1931, give it just a short mention (“since we shall not make anyuse of the Fredholm formulas later on”); while Smithies (1958, Chap. V) and Hochstadt (1973, Chap. 7),acknowledging the fact that “classical” Fredholm theory yields a number of results that functional ana-lytic techniques do not, postpone Fredholm’s theory to a later chapter; whereas Baker (1977), Porter andStirling (1990), Prössdorf and Silbermann (1991), Hackbusch (1995), and Kress (1999) ignore the Fredholmdeterminant altogether. Among the newer books on linear integral equations, the monumental four vol-ume work of Fenyo and Stolle (1982–1984) is one of the few we know of that give Fredholm determinantsa balanced treatment.

4Though we can only speculate about the reasons, there is something like a disapproving attitudetowards determinants in general that seems to be quite common among people working in “continuous”applied mathematics. Here are a few scattered examples: Meyer (2000) writes at the beginning of thechapter on determinants (p. 460): “Today matrix and linear algebra are in the main stream of appliedmathematics, while the role of determinants has been relegated to a minor backwater position.” Axler(1995) has a paper with the provocative title “Down with Determinants!” and a correspondingly workedout textbook on linear algebra (1997). The quintessential book of Golub and Van Loan (1996) on “MatrixComputations” does not explicitly address the computation of determinants at all, it is only implicitlystated as part of Theorem 3.2.1. Higham (2002) writes at the beginning of Section 14.6: “Like the matrixinverse, the determinant is a quantity that rarely needs to be computed.” He then continues with theargument, well known to every numerical analyst, that the determinant cannot be used as a measureof ill conditioning since it scales as det(αA) = αm det(A) for a m× m-matrix A, α ∈ R. Certainly there

Page 3: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3

if either the eigenvalues of the integral operator are, more or less, explicitly knownor if an alternative analytic expression has been found that is numerically moreaccessible—in each specific case anew, lacking a general procedure.

The Nyström-type method advocated in this paper. In contrast, we study a simplegeneral numerical method for Fredholm determinants which is exceptionally effi-cient for smooth kernels, yielding small absolute errors (i.e., errors that are smallwith respect to the scale det(I) = 1 inherently given by the operator determinant).To this end we follow the line of thought of Nyström’s (1930) classical quadraturemethod for the numerical solution of the Fredholm equation (1.1). Namely, given aquadrature rule

Q( f ) =m

∑j=1

wj f (xj) ≈∫ b

af (x) dx,

Nyström discretized (1.1) as the linear system

ui + zm

∑j=1

wjK(xi, xj)uj = f (xi) (i = 1, . . . , m), (1.3)

which has to be solved for ui ≈ u(xi) (i = 1, . . . , m). Nyström’s method is extremelysimple and, yet, extremely effective for smooth kernels. So much so that Delves andMohamed (1985, p. 245), in a chapter comparing different numerical methods forFredholm equations of the second kind, write:

Despite the theoretical and practical care lavished on the more com-plicated algorithms, the clear winner of this contest has been theNyström routine with the m-point Gauss–Legendre rule. This rou-tine is extremely simple; it includes a call to a routine which pro-vides the points and weights for the quadrature rule, about twelvelines of code to set up the Nyström equations and a call to the rou-tine which solves these equations. Such results are enough to makea numerical analyst weep.

By keeping this conceptual and algorithmic simplicity, the method studied in thispaper approximates the Fredholm determinant d(z) simply by the determinant ofthe m×m-matrix that is applied to the vector (ui) in the Nyström equation (1.3):

dQ(z) = det(δij + z wiK(xi, xj)

)mi,j=1 . (1.4)

If the weights wj of the quadrature rule are positive (which is always the betterchoice), we will use the equivalent symmetric variant

dQ(z) = det(

δij + z w1/2i K(xi, xj)w1/2

j

)m

i,j=1. (1.5)

is much truth in all of their theses, and Cramer’s rule and the characteristic polynomial, which werethe most common reasons for a call to the numerical evaluation of determinants (Stewart 1998, p. 176),have most righteously been banned from the toolbox of a numerical analyst for reasons of efficiency.However, with respect to the infinite dimensional case, the elimination of determinants from the thinkingof numerical analysts as a subject of computations might have been all too successful. For instance, thescaling argument does not apply in the infinite dimensional case: operator determinants det(I + A) aredefined for compact perturbations of the identity, which perfectly determines the scaling since, for α 6= 1,α(I + A) cannot be written in the form I + A with another compact operator A. (This is because theidentity operator is not compact then.)

Page 4: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

4 F. BORNEMANN

Using Gauss–Legendre or Curtis–Clenshaw quadrature rules, the computationalcost5 of the method is of order O(m3). The implementation in Matlab, or Mathe-matica, is straightforward and takes just a few lines of code.6 In Matlab:

function d = DetNyström(K,z,a,b,m)

[w,x] = QuadratureRule(a,b,m);

w = sqrt(w);

[xj,xi] = meshgrid(x,x);

d = det(eye(m)+z*(w'*w).*K(xi,xj));

In Mathematica:

ü Clenshaw-Curtis Quadrature (following J. Waldvogel, BIT 43, pp. 1-8, 2003)

ClenshawCurtis@8a_, b_<, m_, dig_D := ModuleB8c = SetPrecision@Cos@π Range@0, mDê mD, digD, M = Range@1, m − 1, 2D,

l, n, v, g, w<, l = Length@MD; n = m − l; v = SetPrecision@82 ê M êHM − 2L, 1ê Take@M, −1D, Table@0, 8n<D< êê Flatten, digD;

v = −Drop@v, −1D − Take@v, 8Length@vD, 2, −1<D; g = Table@−1, 8m<D;g@@l + 1DD += m; g@@n + 1DD += m; g ê= Im2 + Mod@m, 2DM;

w = Re@InverseFourier@g + v, FourierParameters → 8−1, 1<DDê m;

w = Append@w, w@@1DDD; :Hb − aL

2 w,

1 − c

2 a +

1 + c

2 b>F

ü Nyström-type Method for the Fredholm Determinant

DetNyström@K_, z_, a_, b_, m_D :=

ModuleB8w, x<, 8w, x< = QuadratureRule@a, b, mD; w = w ;

Det@IdentityMatrix@mD + z Outer@Times, w, wD Outer@K, x, xDDF;

ü Example for the Sine-Kernel

QuadratureRule@a_, b_, m_D := ClenshawCurtis@8a, b<, m − 1, 20D

K@x_, y_D := Sinc@π Hx − yLD;

N@DetNyström@K, −1, 0, 1 ê10, 10D, 16D

0.9000272717982592

Strictly speaking we are not the first to suggest this simple method, though.In fact, it was Hilbert himself (1904, pp. 52–60) in his very first paper on integralequations, who motivated7 the expression (1.2) of the Fredholm determinant by es-sentially using this method with the rectangular rule for quadrature, proving locallyuniform convergence; see also Hilbert (1912, pp. 4–10) and, for the motivationalargument given just heuristically, without a proof of convergence, Whittaker andWatson (1927, Sect. 11.2), Tricomi (1957, pp. 66–68) (who speaks of a “poetic license”to be applied “without too many scruples”), Smithies (1958, pp. 65–68), Hochstadt(1973, pp. 243–239), and Fenyo and Stolle (1982–1984, Vol. II, pp. 82–84)—to namejust a few but influential cases. Quite astonishingly, despite of all its presence asa motivational tool in the expositions of the classical theory, we have found justone example of the use of this method (with Gauss–Legendre quadrature) in an ac-tual numerical calculation: a paper by the physicists Reinhardt and Szabo (1970) onlow-energy elastic scattering of electrons from hydrogen atoms. However, the errorestimates (Theorem 6.2) that we will give in this paper seem to be new at least; wewill prove that the approximation error essentially behaves like the quadrature errorfor the sections x 7→ K(x, y) and y 7→ K(x, y) of the kernel. In particular, we will getexponential convergence rates for analytic kernels.

Examples. Perhaps the generality and efficiency offered by our direct numericalapproach to Fredholm determinants, as compared to analytic methods if they are

5The computational cost of O(m3) for the matrix determinant using either Gaussian elimination withpartial pivoting (Stewart 1998, p. 176) or Hyman’s method (Higham 2002, Sect. 14.6) clearly dominatesthe cost of O(m log m) for the weights and points of Clenshaw–Curtis quadrature using the FFT, as well asthe cost of O(m2) for Gauss–Legendre quadrature using the Golub–Welsh algorithm; for implementationdetails of these quadrature rules see Waldvogel (2006), Trefethen (2008), and the appendix of this paper.The latter paper carefully compares Clenshaw–Curtis with Gauss–Legendre and concludes (p. 84): “Gaussquadrature is a beautiful and powerful idea. Yet the Clenshaw–Curtis formula has essentially the sameperformance for most integrands.”

6The command [w,x] = QuadratureRule(a,b,m) is supposed to supply the weights and points ofa m-point quadrature rule on the interval [a, b] as a 1×m vector w and a m× 1-vector x, respectively. ForGauss–Legendre and Clenshaw–Curtis, such a Matlab code can be found in the appendix.

7Fredholm himself does not give the slightest hint of a motivation in his early papers (1900, 1903). Heapparently conceived his determinant in analogy to similar expressions that his fellow countryman vonKoch (1892) had obtained for infinite matrices a few years earlier; see Fredholm (1909, p. 95), Dieudonné(1981, p. 99), or Pietsch (2007, p. 409).

Page 5: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 5

available at all, is best appreciated by an example. The probability E2(0; s) that aninterval of length s does not contain, in the bulk scaling limit of level spacing 1,an eigenvalue of the Gaussian unitary ensemble (GUE) is given by the Fredholmdeterminant of the sine kernel,

E2(0; s) = det(

I − As�L2(0,s)

), Asu(x) =

∫ s

0

sin(π(x− y))π(x− y)

u(y) dy ;

see Gaudin (1961) and Mehta (2004, Sect. 6.3). Gaudin has further shown that theeigenfunctions of this selfadjoint integral operator are exactly given by a particularfamily of special functions, namely the radial prolate spheroidal wave functions withcertain parameters. Using tables (Stratton, Morse, Chu, Little and Corbató 1956) ofthese functions he was finally able to evaluate E2(0; s) numerically. On the otherhand, in an admirably intricate analytic tour de force Jimbo, Miwa, Môri and Sato(1980) expressed the Fredholm determinant of the sine kernel as

Es(0; s) = exp(∫ πs

0

σ(x)x

dx)

(1.6)

in terms of the sigma, or Hirota, representation of the fifth Painlevé equation, namely

(xσ′′)2 + 4(xσ′ − σ)(xσ′ − σ + σ′2) = 0, σ(x) ∼ − xπ− x2

π2 (x → 0),

see also Mehta (2004, Chap. 21) and Tracy and Widom (2000, Sect. 4.1). With respectto the numerical evaluation, the latter two authors conclude in a footnote, mostprobably by comparing to Gaudin’s method: “Without the Painlevé representations,the numerical evaluation of the Fredholm determinants is quite involved.” However,one does not need to know more than the smooth kernel sin(π(x− y))/(π(x− y))ifself to approximate E2(0; s) with the method of this paper. For instance, the Gauss–Legendre rule with just m = 5 quadrature points already gives, in 0.2 ms computingtime, 15 accurate digits of the value

E2(0, 0.1) = 0.90002 72717 98259 · · · ,

that is, by calculating the determinant of a 5× 5-matrix easily built from the kernel.Even though it is satisfying to have an alternative and simpler way of calculat-

ing already known quantities, it is far more exciting to be able to calculate quantitiesthat otherwise have defeated numerical evaluations so far. For instance, the joint dis-tribution functions of the Airy and the Airy1 processes are given as determinantsof systems of integral operators, see Prähofer and Spohn (2002), Johansson (2003),Sasamoto (2005) and Borodin et al. (2007). Even though a nonlinear partial differ-ential equation of third order in three variables has been found by Adler and vanMoerbeke (2005, Eq. (4.12)) for the logarithm of the joint distribution function ofthe Airy process at two different times, this masterful analytic result is probably ofnext to no numerical use. And in any case, no such analytic results are yet knownfor the Airy1 process. However, the Nyström-type method studied in this paper caneasily be extended to systems of integral operators. In this way, we have succeededin evaluating the two-point correlation functions of both stochastic processes, seeSection 8.

Page 6: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

6 F. BORNEMANN

Outline of this paper. For the proper functional analytic setting, in Section 2 wereview some basic facts about trace class and Hilbert–Schmidt operators. In Section 3

we review the concept of the determinant det(I + A) for trace class operators A andits relation to the Fredholm determinant. In Section 4 we study perturbation boundsimplying that numerical calculations of determinants can only be expected to be ac-curate with respect to absolute errors in general. In Section 5 we use the functionalanalytic formulation of the problem to obtain convergence estimates for projectionmethods of Galerkin and Ritz–Galerkin type. The convergence rate is shown to de-pend on a proper balance between the decay of the singular values of the operatorand the growth of bounds on the derivatives of the corresponding singular func-tions. This is in sharp contrast with Section 6, where we study the convergence ofthe Nyström-type method (1.5) by directly addressing the original definition of theFredholm determinant. Here, only the smoothness properties of the kernel enter theconvergence estimates. It turns out that, for kernels of low regularity, the order ofconvergence of the Nyström-type method can be even higher than that of a Ritz–Galerkin method. In Section 7 we give examples for the exponential convergencerates enjoyed by analytic kernels. To this end we discuss the details of the numericalevaluation of the determinants of the sine and Airy kernels, which express the prob-ability distributions E2(0; s) and F2(s) (the Tracy–Widom distribution) of randommatrix theory. Finally, after extending the Nyström-type method to systems of inte-gral operators we report in Section 8 on the numerical evaluation of the two-pointcorrelation functions of the Airy and Airy1 processes.

2. Trace Class and Hilbert–Schmidt Operators. We begin by recalling somebasic material about the spectral theory of nonselfadjoint compact operators, whichcan be found, e.g., in Gohberg, Goldberg and Kaashoek (1990), Lax (2002) and Simon(2005). We consider a complex, separable Hilbert spaceH with an inner product 〈·, · 〉that is linear in the second factor and conjugate linear in the first. The set of boundedlinear operators will be denoted by B(H), the compact operators by J∞(H). Thespectrum of a compact operator A ∈ J∞(H) has no non-zero limit point; its non-zero points are eigenvalues of finite algebraic multiplicity. We list these eigenvaluesas (λn(A))N(A)

n=1 , counting multiplicity, where N(A) is either a finite non-negativeinteger or infinity, and order them by

|λ1(A)| > |λ2(A)| > · · · .

The positive eigenvalues

s1(A) > s2(A) > · · · > 0

of the associated positive-semidefinite, selfadjoint operator

|A| = (A∗A)1/2

are called the singular values of A. Correspondingly, there is the Schmidt or singular-value representation of A, that is, the norm convergent expansion

A =N(|A|)

∑n=1

sn(A)〈un, · 〉vn, (2.1)

Page 7: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 7

where the un and vn are certain (not necessarily complete) orthonormal sets in H.Note that sn(A) = |λn(A)| if A is selfadjoint. In general we have Weyl’s inequality

N

∑n=1|λn(A)|p 6

N

∑n=1

sn(A)p (N 6 N(A), 1 6 p < ∞). (2.2)

The Schatten–von Neumann classes of compact operators are defined as

Jp(H) = {A ∈ J∞(H) :N(|A|)

∑n=1

sn(A)p < ∞} (1 6 p < ∞)

with the corresponding norm8

‖A‖Jp =

(N(|A|)

∑n=1

sn(A)p

)1/p

.

The operator norm on J∞(H) perfectly fits into this setting if we realize that

‖A‖ = s1(A) = maxn=1,...,N(|A|)

sn(A) = ‖A‖J∞ .

There are the continuous embeddings Jp(H) ⊂ Jq(H) for 1 6 p 6 q 6 ∞ with

‖A‖Jq 6 ‖A‖Jp .

The classes Jp(H) are two-sided operator ideals in B(H), that is, for A ∈ Jp(H)(1 6 p 6 ∞) and B ∈ B(H) we have AB, BA ∈ Jp(H) with

‖AB‖Jp 6 ‖A‖Jp‖B‖, ‖BA‖Jp 6 ‖B‖ ‖A‖Jp . (2.3)

Of special interest to us are the trace class operators J1(H) and the Hilbert–Schmidtoperators J2(H). The product of two Hilbert–Schmidt operators is of trace class:

‖AB‖J1 6 ‖A‖J2‖B‖J2 (A, B ∈ J2(H)).

The trace of a trace class operator A is defined by

tr(A) =∞

∑n=1〈un, Aun〉

for any orthonormal basis (un)n. A deep theorem of Lidskii’s (Simon 2005, Chap. 3)tells us that

tr(A) =N(A)

∑n=1

λn(A), (2.4)

which implies by Weyl’s inequality (2.2) that

|tr(A)| 6N(A)

∑n=1|λn(A)| 6 tr(|A|) = ‖A‖J1 . (2.5)

8In matrix theory these norms are not commonly used—with the exception of p = 2: ‖A‖J2 is thenthe Schur or Frobenius norm of the matrix A.

Page 8: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

8 F. BORNEMANN

Likewise, for a Hilbert–Schmidt operator A ∈ J2(H) we have

tr(A2) =N(A)

∑n=1

λn(A)2, |tr(A2)| 6N(A)

∑n=1|λn(A)|2 6 ‖A‖2

J2. (2.6)

Integral operators with L2-kernel. In the Hilbert space H = L2(a, b) of square-integrable functions on a finite interval (a, b) the Hilbert–Schmidt operators are ex-actly given by the integral operators with L2-kernel. That is, there is a one-to-onecorrespondence (Simon 2005, Thm. 2.11) between A ∈ J2(H) and K ∈ L2((a, b)2)mediated through

Au(x) =∫ b

aK(x, y)u(y) dy (2.7)

with equality of norms ‖A‖J2 = ‖K‖L2 : the spaces J2(H) and L2((a, b)2) are thusisometrically isomorphic. In particular, by (2.6) and a well known basic result oninfinite products (Knopp 1964, p. 232), we get for such operators that

N(A)

∏n=1

(1 + λn(A)) converges (absolutely) ⇔N(A)

∑n=1

λn(A) converges (absolutely).

Since the product is a natural candidate for the definition of det(I + A) it makessense requiring A to be of trace class; for then, by (2.5), the absolute convergence ofthe sum can be guaranteed.

Integral operators with a continuous kernel. A continuous kernel K ∈ C([a, b]2) iscertainly square-integrable. Therefore, the induced integral operator (2.7) defines aHilbert–Schmidt operator A on the Hilbert space H = L2(a, b). Moreover, other thanfor L2 kernels in general, the integral∫ b

aK(x, x) dx

over the diagonal of (a, b)2 is now well defined and constitutes, in analogy to thematrix case, a “natural” candidate for the trace of the integral operator. Indeed, if anintegral operator A with continuous kernel is of trace class, one can prove (Gohberget al. 2000, Thm. 8.1)

tr(A) =∫ b

aK(x, x) dx. (2.8)

Unfortunately, however, just the continuity of the kernel K does not guarantee theinduced integral operator A to be of trace class.9 Yet, there is some encouragingpositive experience stated by Simon (2005, p. 25):

However, the counter-examples which prevent nice theorems fromholding are generally rather contrived so that I have found the fol-lowing to be true: If an integral operator with kernel K occurs insome ‘natural’ way and

∫|K(x, x)| dx < ∞, then the operator can

(almost always) be proven to be trace class (although sometimesonly after some considerable effort).

9A counter-example was discovered by Carleman (1918), see also Gohberg et al. (2000, p. 71).

Page 9: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 9

Nevertheless, we will state some simple criteria that often work well:

1. If the continuous kernel K can be represented in the form

K(x, y) =∫ d

cK1(x, y)K2(z, y) dz (x, y ∈ [a, b])

with K1 ∈ L2((a, b) × (c, d)), K2 ∈ L2((c, d) × (a, b)), then the induced in-tegral operator A is trace class on L2(a, b). This is, because A can then bewritten as the product of two Hilbert–Schmidt operators.

2. If K(x, y) and ∂yK(x, y) are continuous on [a, b]2, then the induced integraloperator A is trace class on L2(a, b). This is, because we can write A bypartial integration in the form

Au(x) = K(x, b)∫ b

au(y) dy−

∫ b

a

(∫ b

y∂zK(x, z) dz

)u(y) dy

as a sum of a rank one operator and an integral operator that is trace classby the first criterion. In particular, integral operators with smooth kernelsare trace class (Lax 2002, p. 345).

3. A continuous Hermitian10 kernel K(x, y) on [a, b] that satisfies a Hölder con-dition in the second argument with exponent α > 1/2, namely

|K(x, y1)− K(x, y2)| 6 C|y1 − y2|α (x, y1, y2 ∈ [a, b]),

induces an integral operator A that is trace class on L2(a, b); see Gohberget al. (2000, Thm. IV.8.2).

4. If the continuous kernel K induces a selfadjoint, positive-semidefinite inte-gral operator A, then A is trace class (Gohberg et al. 2000, Thm. IV.8.3). Thehypothesis on A is fulfilled for positive-semidefinite kernels K, that is, if

n

∑j,k=1

zjzkK(xj, xk) > 0 (2.9)

for any x1, . . . , xn ∈ (a, b), z ∈ Cn and any n ∈N (Simon 2005, p. 24).

3. Definition and Properties of Fredholm and Operator Determinants. In thissection we give a general operator theoretical definition of infinite dimensional de-terminants and study their relation to the Fredholm determinant. For a trace classoperator A ∈ J1(H) there are several equivalent constructions that all define oneand the same entire function

d(z) = det(I + zA) (z ∈ C);

in fact, each construction has been chosen at least once, in different places of theliterature, as the basic definition of the operator determinant:

10An L2-kernel K is Hermitian if K(x, y) = K(y, x). This property is equivalent to the fact that theinduced Hilbert–Schmidt integral operator A is selfadjoint, A∗ = A.

Page 10: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

10 F. BORNEMANN

1. Gohberg and Kreın (1969, p. 157) define the determinant by the locally uni-formly convergent (infinite) product

det(I + zA) =N(A)

∏n=1

(1 + zλn(A)), (3.1)

which possesses zeros exactly at zn = −1/λn(A), counting multiplicity.2. Gohberg et al. (1990, p. 115) define the determinant as follows. Given any

sequence of finite rank operators An with An → A converging in trace classnorm, the sequence of finite dimensional determinants11

det(

I + zAn�ran(An)

)(3.2)

(which are polynomials in z) converges locally uniform to det(I + zA), in-dependently of the choice of the sequence An. The existence of at least onesuch sequence follows from the singular value representation (2.1).

3. Dunford and Schwartz (1963, p. 1029) define the determinant by what isoften called Plemelj’s formula12

det(I + zA) = exp(tr log(I + zA)) = exp

(−

∑n=1

(−z)n

ntrAn

), (3.3)

which converges for |z| < 1/|λ1(A)| and can analytically be continued asan entire function to all z ∈ C.

4. Grothendieck (1956, p. 347) and Simon (1977, p. 254) define the determinantmost elegantly with a little exterior algebra (Greub 1967). With

∧n(A) ∈J1(

∧n(H)) being the nth exterior product of A, the power series

det(I + zA) =∞

∑n=0

zntr∧n

(A) (3.4)

converges for all z ∈ C. Note that tr∧n(A) = ∑i1<···<in λi1(A) · · · λin(A) is

just the nth symmetric function of the eigenvalues of A.Proofs of the equivalence can be found in (Gohberg et al. 2000, Chap. 2) and (Simon2005, Chap. 3). We will make use of all of them in the course of this paper. We statetwo important properties (Simon 2005, Thm. 3.5) of the operator determinant: Firstits multiplication formula,

det(I + A + B + AB) = det(I + B) det(I + A) (A, B ∈ J1(H)), (3.5)

then the characterization of invertibility: det(I + A) 6= 0 if and only if the inverseoperator (I + A)−1 exists.

11Gohberg et al. (2000) have later extended this idea to generally define traces and determinants onembedded algebras of compact operators by a continuous extension from the finite dimensional case.Even within this general theory the trace class operators enjoy a most unique position: it is only for themthat the values of trace and determinant are independent of the algebra chosen for their definition. On thecontrary, if A is Hilbert–Schmidt but not trace class, by varying the embedded algebra, the values of thetrace tr(A) can be given any complex number and the values of the determinant det(I + A) are eitheralways zero or can be made to take any value in the set C \ {0} (Gohberg et al. 2000, Chap. VII).

12Plemelj (1904, Eq. (62)) had given a corresponding form of the Fredholm determinant for integraloperators. However, it can already be found in Fredholm (1903, p. 384).

Page 11: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 11

The matrix case. In Section 6 we will study the convergence of finite dimensionaldeterminants to operator determinants in terms of the power series (3.4). Therefore,we give this series a more common look and feel for the case of a matrix A ∈ Cm×m.By evaluating the traces with respect to a Schur basis of A one gets

tr∧n

(A) = ∑i1<···<in

det(Aip ,iq)np,q=1 =

1n!

m

∑i1,...,in=1

det(Aip ,iq)np,q=1,

that is, the sum of all n× n principal minors of A. The yields the von Koch (1892)form of the matrix determinant

det(I + zA) =∞

∑n=0

zn

n!

m

∑i1,...,in=1

det(Aip ,iq)np,q=1 (A ∈ Cm×m). (3.6)

(In fact, the series must terminate at n = m since det(I + zA) is a polynomial of de-gree m in z.) A more elementary proof of this classical formula, by a Taylor expansionof the polynomial det(I + zA), can be found, e.g., in Meyer (2000, p. 495).

The Fredholm determinant for integral operators with continuous kernel. Suppose thatthe continuous kernel K ∈ C([a, b]2) induces an integral operator A that is trace classon the Hilbert spaceH = L2(a, b). Then, the traces of

∧n(A) evaluate to (Simon 2005,Thm. 3.10)

tr∧n

(A) =1n!

∫(a,b)n

det(K(tp, tq))np,q=1 dt1 · · · dtn (n = 0, 1, 2, . . .).

The power series representation (3.4) of the operator determinant is therefore exactlyFredholm’s expression (1.2), that is,

det(I + zA) =∞

∑n=0

zn

n!

∫(a,b)n

det(K(tp, tq))np,q=1 dt1 · · · dtn. (3.7)

The similarity with von Koch’s formula (3.6) is striking and, in fact, it was just ananalogy in form that had led Fredholm to conceive his expression for the deter-minant. It is important to note, however, that the right hand side of (3.7) perfectlymakes sense for any continuous kernel, independently of whether the correspondingintegral operator is trace class or not.

The regularized determinant for Hilbert–Schmidt operators. For a general Hilbert–Schmidt operator we only know the convergence of ∑n λ(A)2 but not of ∑n λn(A).Therefore, the product (3.1), which is meant to define det(I + zA), is not knownto converge in general. Instead, Hilbert (1904) and Carleman (1921) introduced theentire function13

det2(I + zA) =N(A)

∏n=1

(1 + zλn(A))e−zλn(A) (A ∈ J2(H)),

13In fact, using such exponential “convergence factors” is a classical technique in complex analysis toconstruct, by means of infinite products, entire functions from their intended sets of zeros, see Ablowitzand Fokas (2003, Sect. 3.6). A famous example is

1Γ(z)

= zeγz∞

∏n=1

(1 +

zn

)e−z/n,

which corresponds to the eigenvalues λn(A) = 1/n of a Hilbert–Schmidt operator that is not trace class.

Page 12: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

12 F. BORNEMANN

which also has the property to possess zeros exactly at zn = −1/λn(A), countingmultiplicity. Plemelj’s formula (3.3) becomes (Gohberg et al. 2000, p. 167)

det2(I + zA) = exp

(−

∑n=2

(−z)n

ntrAn

)

for |z| < 1/|λ1(A)|, which perfectly makes sense since A2, A3, . . . are trace class forA being Hilbert–Schmidt.14 Note that for trace class operators we have

det(I + zA) = det2(I + zA) exp(z · trA) (A ∈ J1(H)).

For integral operators A of the form (2.7) with a continuous kernel K onH = L2(a, b)the Fredholm determinant (1.2) is related to the Hilbert–Carleman determinant bythe equation (Hochstadt 1973, p. 250)

d(z) = det2(I + zA) exp(z∫ b

aK(x, x) dx)

in general, even if A is not of trace class. It is important to note, though, that ifA 6∈ J1(H) with such a kernel, we have

∫ ba K(x, x) dx 6= tr(A) simply because tr(A)

is not well defined by (2.4) anymore then.

4. Perturbation Bounds. In studying the conditioning of operator and matrixdeterminants we rely on the fundamental perturbation estimate

|det(I + A)− det(I + B)| 6 ‖A− B‖J1 exp(1 + max(‖A‖J1 , ‖B‖J1)

)(4.1)

for trace class operators, which can beautifully be proven by means of complex anal-ysis (Simon 2005, p. 45). This estimate can be put to the form

|det(I + (A + E))− det(I + A)| 6 e1+‖A‖J1 · ‖E‖J1 + O(‖E‖2J1

), (4.2)

showing that the condition number κabs of the operator determinant det(I + A), withrespect to absolute errors measured in trace class norm, is bounded by

κabs 6 e1+‖A‖J1 .

This bound can considerably be improved for certain selfadjoint operators that willplay an important role in Section 7.

14Equivalently one can define (Simon 2005, p. 50) the regularized determinant by

det2(I + zA) = det(I + ((I + zA)e−zA − I)) (A ∈ J2(H)).

This is because one can then show (I + zA)e−zA − I ∈ J1(H). For integral operators A on L2(a, b) withan L2 kernel K, Hilbert (1904, p. 82) had found the equivalent expression

det2(I + zA) =∞

∑n=0

zn

n!

∫(a,b)n

∣∣∣∣∣∣∣∣∣∣∣

0 K(t1, t2) · · · K(t1, tn)K(t2, t1) 0 · · · K(t2, tn)

......

. . ....

K(tn, t1) K(tn, t2) · · · 0

∣∣∣∣∣∣∣∣∣∣∣dt1 · · · dtn,

simply replacing the problematic “diagonal” entries K(tj, tj) in Fredholm’s determinant (3.7) by zero.Simon (2005, Thm. 9.4) gives an elegant proof of Hilbert’s formula.

Page 13: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 13

Lemma 4.1. Let A ∈ J1(H) be selfadjoint, positive-semidefinite with λ1(A) < 1. If‖E‖J1 < ‖(I − A)−1‖−1 then

|det(I − (A + E))− det(I − A)| 6 ‖E‖J1 . (4.3)

That is, the condition number κabs of the determinant det(I − A), with respect to absoluteerrors measured in trace class norm, is bounded by κabs 6 1.

Proof. Because of 1 > λ1(A) > λ2(A) > · · · > 0 there exists the inverse operator(I− A)−1. The product formula (3.1) implies det(I− A) > 0, the multiplicativity (3.5)of the determinant gives

det(I − (A + E)) = det(I − A) det(I − (I − A)−1E).

Upon applying Plemelj’s formula (3.3) and the estimates (2.3) and (2.5) we get

| log det(I − (I − A)−1E)| =∣∣∣∣∣tr(

∑n=1

((I − A)−1E)n

n

)∣∣∣∣∣6

∑n=1

‖(I − A)−1‖n · ‖E‖nJ1

n= log

(1

1− ‖(I − A)−1‖ · ‖E‖J1

)if ‖(I − A)−1‖ · ‖E‖J1 < 1. Hence, exponentiation yields

1− ‖(I − A)−1‖ · ‖E‖J1 6 det(I − (I − A)−1E)

61

1− ‖(I − A)−1‖ · ‖E‖J1

6 1 + ‖(I − A)−1‖ · ‖E‖J1 ,

that is

|det(I − (A + E))− det(I − A)| 6 det(I − A) · ‖(I − A)−1‖ · ‖E‖J1 .

Now, by the spectral theorem for bounded selfadjoint operators we have

‖(I − A)−1‖ =1

1− λ1(A)6

N(A)

∏n=1

11− λn(A)

=1

det(I − A)

and therefore det(I − A) · ‖(I − A)−1‖ 6 1, which finally proves the assertion.

Thus, for the operators that satisfy the assumptions of this lemma the determi-nant is a really well conditioned quantity—with respect to absolute errors, like theeigenvalues of a Hermitian matrix (Golub and Van Loan 1996, p. 396).

Implications on the accuracy of numerical methods. The Nyström-type method ofSection 6 requires the calculation of the determinant det(I + A) of some matrix A ∈Cm×m. In the presence of roundoff errors, a backward stable method like Gaussianelimination with partial pivoting (Higham 2002, Sect. 14.6) gives a result that is exactfor some matrix A = A + E with

|Ej,k| 6 ε|Aj,k| (j, k = 1, . . . , m) (4.4)

Page 14: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

14 F. BORNEMANN

where ε is a small multiple of the unit roundoff error of the floating-point arithmeticused. We now use the perturbation bounds of this section to estimate the resultingerror in the value of determinant. Since the trace class norm is not a monotone matrixnorm (Higham 2002, Def. 6.1), we cannot make direct use of the componentwiseestimate (4.4). Instead, we majorize the trace class norm of m×m matrices A by theHilbert–Schmidt (Frobenius) norm, which is monotone, using

‖A‖J1 6√

m‖A‖J2 , ‖A‖J2 =

(m

∑j,k=1|Aj,k|2

)1/2

.

Thus, the general perturbation bound (4.2) yields the following a priori estimate ofthe roundoff error affecting the value of the determinant:

|det(I + A)− det(I + A)| 6√

m‖A‖J2 exp(1 + ‖A‖J1

)· ε + O(ε2).

If the matrix A satisfies the assumptions of Lemma 4.1, the perturbation bound (4.3)gives the even sharper estimate

|det(I − A)− det(I − A)| 6√

m‖A‖J2 · ε. (4.5)

Therefore, if det(I − A) � ‖A‖J2 , we have to be prepared that we probably cannotcompute the determinant to the full precision given by the computer arithmetic used.Some digits will be lost. A conservative estimate stemming from (4.5) predicts theloss of at most

log10

(√m · ‖A‖J2

det(I − A)

)decimal places. For instance, this will affect the tails of the probability distributionsto be calculated in Section 7.

5. Projection Methods. The general idea (3.2) of defining the infinite dimen-sional determinant det(I + A) for a trace class operator A by a continuous exten-sion from the finite dimensional case immediately leads to the concept of a projec-tion method of Galerkin type. We consider a sequence of m-dimensional subspacesVm ⊂ H together with their corresponding orthonormal projections

Pm : H → Vm.

The Galerkin projection Pm APm of the trace class operator A is of finite rank. Givenan orthonormal basis φ1, . . . , φm of Vm, its determinant can be effectively calculatedas the finite dimensional expression

det(I + z Pm APm) = det (I + z Pm APm�Vm) = det(δij + z 〈φi, Aφj〉

)mi,j=1 (5.1)

if the matrix elements 〈φi, Aφj〉 are numerically accessible.Because of ‖Pm‖ 6 1, and thus ‖Pm APm‖J1 6 ‖A‖J1 6 1, the perturbation

bound (4.1) gives the simple error estimate

|det(I + z Pm APm)− det(I + z A)| 6 ‖Pm APm − A‖J1 · |z| e1+|z|·‖A‖J1 . (5.2)

For the method to be convergent we therefore have to show that Pm APm → A intrace class norm. By a general result about the approximation of trace class operators(Gohberg et al. 1990, Thm. 4.3) all we need to know is that Pm converges pointwise15

15In infinite dimensions Pm cannot converge in norm since the identity operator is not compact.

Page 15: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 15

to the identity operator I. This pointwise convergence is obviously equivalent to theconsistency of the family of subspaces Vm, that is,

∞⋃m=1

Vm is a dense subspace of H. (5.3)

In summary, we have proven the following theorem.Theorem 5.1. Let A be a trace class operator. If the sequence of subspaces satisfies the

consistency condition (5.3), the corresponding Galerkin approximation (5.1) of the operatordeterminant converges,

det(I + z Pm APm)→ det(I + z A) (m→ ∞),

uniformly for bounded z.A quantitative estimate of the error, that is, in view of (5.2), of the projection

error ‖Pm APm − A‖J1 in trace class norm, can be based on the singular value rep-resentation (2.1) of A and its finite-rank truncation AN : (We assume that A is non-degenerate, that is, N(|A|) = ∞; since otherwise we could simplify the following byputting AN = A.)

A =∞

∑n=1

sn(A)〈un, · 〉vn, AN =N

∑n=1

sn(A)〈un, · 〉vn.

We obtain, by using ‖Pm‖ 6 1 once more,

‖Pm APm − A‖J1 6 ‖Pm APm − Pm AN Pm‖J1 + ‖Pm AN Pm − AN‖J1 + ‖AN − A‖J1

6 2‖AN − A‖J1 + ‖Pm AN Pm − AN‖J1

6 2∞

∑n=N+1

sn(A) +N

∑n=1

sn(A)‖〈Pmun, · 〉Pmvn − 〈un, · 〉vn‖J1

6 2∞

∑n=N+1

sn(A) +N

∑n=1

sn(A) (‖un − Pmun‖+ ‖vn − Pmvn‖) . (5.4)

There are two competing effects that contribute to making this error bound small:First, there is the convergence of the truncated series of singular values,

∑n=N+1

sn(A)→ 0 (N → ∞),

which is independent of m. Second, there is, for fixed N, the collective approximation

Pmun → un, Pmvn → vn (m→ ∞)

of the first N singular functions un, vn (n = 1, . . . , N). For instance, given ε > 0, wecan first choose N large enough to push the first error term in (5.4) below ε/2. Afterfixing such an N, the second error term can be pushed below ε/2 for m large enough.This way we have proven Theorem 5.1 once more. However, in general the conver-gence of the second term might considerably slow down for growing N. Therefore,

Page 16: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

16 F. BORNEMANN

a good quantitative bound requires a carefully balanced choice of N (depending onm), which in turn requires some detailed knowledge about the decay of the singularvalues on the one hand and of the growth of the derivatives of the singular functionson the other hand (see the example at the end of this section). While some general re-sults are available in the literature for the singular values—e.g., for integral operatorsA induced by a kernel K ∈ Ck−1,1([a, b]2) the bound

sn(A) = O(n−k− 12 ) (n→ ∞) (5.5)

obtained by Smithies (1937, Thm. 12)—the results are sparse for the singular func-tions (Fenyo and Stolle 1982–1984, §8.10). Since the quadrature method in Section 6

does not require any such knowledge, we refrain from stating a general result andcontent ourselves with the case that there is no projection error in the singular func-tions; that is, we consider projection methods of Ritz–Galerkin type for selfadjointoperators.

Theorem 5.2. Let A be a selfadjoint integral operator that is induced by a continuousHermitian kernel K and that is trace class on the Hilbert spaceH = L2(a, b). Assume that Ais not of finite rank and let (un) be an orthonormal basis of eigenfunctions of A. We considerthe associated Ritz projection Pm, that is, the orthonormal projection

Pm : H → Vm = span{u1, . . . , um}.

Note that in this case

det(I + z Pm APm) =m

∏n=1

(1 + zλn(A)).

If K ∈ Ck−1,1([a, b]2), then there holds the error estimate (5.2) with

‖Pm APm − A‖J1 = o(m12−k) (m→ ∞).

If K is bounded analytic on Eρ × Eρ (with the ellipse Eρ defined in Theorem A.2), then theerror estimate improves to

‖Pm APm − A‖J1 = O(ρ−m(1−ε)/4) (m→ ∞),

for any fixed choice of ε > 0.Proof. With the spectral decompositions

A =∞

∑n=1

λn(A)〈un, · 〉un, Pm APm = Am =m

∑n=1

λn(A)〈un, · 〉un,

at hand the bound (5.4) simplifies, for N = m, to

‖Pm APm − A‖J1 =∞

∑n=m+1

|λn(A)|.

Now, by some results of Hille and Tamarkin (1931, Thm. 7.2 and 10.1), we have, forK ∈ Ck−1,1([a, b]2), the eigenvalue decay

λn(A) = o(n−k− 12 ) (n→ ∞) (5.6)

Page 17: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 17

(which is just slightly stronger than Smithies’ singular value bound (5.5)) and, for Kbounded analytic on Eρ × Eρ,

λn(A) = O(ρ−n(1−ε)/4) (n→ ∞);

which proves both assertions.However, for kernels of low regularity, by taking into account the specifics of

a particular example one often gets better results than stated in this theorem. (Anexample with an analytic kernel, enjoying the excellent convergence rates of thesecond part this theorem, can be found in Section 7.)

An example: Poisson’s equation. For a given f ∈ L2(0, 1) the Poisson equation

−u′′(x) = f (x), u(0) = u(1) = 0,

with Dirichlet boundary conditions is solved (Hochstadt 1973, p. 5) by the applica-tion of the integral operator A,

u(x) = A f (x) =∫ 1

0K(x, y) f (y) dy, (5.7)

which is induced by the Green’s kernel

K(x, y) =

{x(1− y) x 6 y,

y(1− x) otherwise.(5.8)

Since K is Lipschitz continuous, Hermitian, and positive definite, we know from theresults of Section 2 that A is a selfadjoint trace class operator on H = L2(0, 1). Theeigenvalues and normalized eigenfunctions of A are those of the Poisson equationwhich are known to be

λn(A) =1

π2n2 , un(x) =√

2 sin(nπx) (n = 1, 2, . . .).

Note that the actual decay of the eigenvalues is better than the general Hille–Ta-markin bound (5.6) which would, because of K ∈ C0,1([0, 1]2), just give λn(A) =o(n−3/2). The trace formulas (2.4) and (2.8) reproduce a classical formula of Euler’s,namely

∑n=1

1π2n2 = tr(A) =

∫ 1

0K(x, x) dx =

∫ 1

0x(1− x) dx =

16

;

whereas, by (3.1) and the product representation of the sine function, the Fredholmdeterminant explicitly evaluates to the entire function

det(I − zA) =∞

∏n=1

(1− z

π2n2

)=

sin(√

z)√z

. (5.9)

The sharper perturbation bound of Lemma 4.1 applies and we get, for each finitedimensional subspace Vm ⊂ H and the corresponding orthonormal projection Pm :H → Vm, the error estimate

|det(I − Pm APm)− det(I − A)| 6 ‖Pm APm − A‖J1 . (5.10)

Now, we study two particular families of subspaces.

Page 18: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

18 F. BORNEMANN

101

102

103

10−4

10−3

10−2

m

|det

(I−

Pm

A P

m)

− d

et(I

−A

)|

Fig. 1. Convergence of Ritz–Galerkin (circles) and Galerkin (dots) for approximating the Fredholm determinantof the integral operator induced by Green’s kernel of Poisson’s equation. The solid line shows the upper bound 1/π2mof Ritz–Galerkin as given in (5.11).

Trigonometric polynomials. Here, we consider the subspaces

Vm = span{sin(nπ ·) : n = 1, . . . , m} = span{un : n = 1, . . . , m},

which are exactly those spanned by the eigenfunctions of A. In this case, the projec-tion method is of Ritz–Galerkin type; the estimates become particularly simple sincewe have the explicit spectral decomposition

Pm APm − A =∞

∑n=m+1

λn(A)〈un, · 〉un

of the error operator. Hence, the error bound (5.10) directly evaluates to

|det(I − Pm APm)− det(I − A)|

6 ‖Pm APm − A‖J1 =∞

∑n=m+1

λn(A) =1

π2

∑n=m+1

1n2 6

1π2m

. (5.11)

Figure 1 shows that this upper bound overestimates the error in the Fredholm deter-minant by just about 20%.

Algebraic polynomials. Here, we consider the subspaces of algebraic polynomialsof order m, that is,

Vm = {u ∈ L2(0, 1) : u is a polynomial of degree at most m− 1}.

We apply the bound given in (5.4) and obtain (keeping in mind that A is selfadjoint)

‖Pm APm − A‖J1 6 2∞

∑n=N+1

λn(A) + 2N

∑n=1

λn(A) ‖un − Pmun‖

with a truncation index N yet to be skilfully chosen. As before in (5.11), the first termof this error bound can be estimated by 2/π2N. For the second term we recall thatthe projection error ‖un − Pmun‖ is, in fact, just the error of polynomial best approxi-mation of the eigenfunction un with respect to the L2-norm. A standard Jackson-type

Page 19: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 19

inequality (DeVore and Lorentz 1993, p. 219) from approximation theory teaches usthat

‖un − Pmun‖ 6ck

mk ‖u(k)n ‖ = ck

(πn)k

mk ,

where ck denotes a constant that depends on the smoothness level k. A fixed eigen-function un (being an entire function in fact) is therefore approximated beyond everyalgebraic order in the dimension m; but with increasingly larger constants for higher“wave numbers” n. We thus get, with some further constant ck depending on k > 2,

‖Pm APm − A‖J1 6 ck

(1N

+Nk−1

(k− 1)mk

).

We now balance the two error terms by minimizing this bound: the optimal trunca-tion index N turns out to be exactly N = m, which finally yields the estimate

|det(I − Pm APm)− det(I − A)| 6 ‖Pm APm − A‖J1 6ck

1− k−1 m−1.

Thus, in contrast to the approximation of a single eigenfunction, for the Fredholmdeterminant the order of the convergence estimate does finally not depend on thechoice of k anymore; we obtain the same O(m−1) behavior as for the Ritz–Galerkinmethod. In fact, a concrete numerical calculation16 shows that this error estimatereally reflects the actual order of the error decay of the Galerkin method, see Figure 1.

Remark. The analysis of this section has shown that the error decay of the pro-jection methods is essentially determined by the decay

∑k=m+1

sk(A)→ 0

of the singular values of A, which in turn is related to the smoothness of the kernelK of the integral operator A. In the next section, the error analysis of Nyström-typequadrature methods will relate in a much more direct fashion to the smoothness of

16By (5.1) all we need to know for implementing the Galerkin method are the matrix elements 〈φi , Aφj〉for the normalized orthogonal polynomials φn (that is, properly rescaled Legendre polynomials) on theinterval [0, 1]. A somewhat lengthy but straightforward calculation shows that these elements are givenby

(〈φi , Aφj〉)i,j =

112 0 b0

0 160 0 b1

b0 0 a1. . .

. . .

b1. . . a2

. . .. . .

with the coefficients

an =1

2(2n + 1)(2n + 5), bn = − 1

4(2n + 3)√

(2n + 1)(2n + 5).

Page 20: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

20 F. BORNEMANN

the kernel K, giving even much better error bounds, a priori and in actual numericalcomputations. For instance, the Green’s kernel (5.7) of low regularity can be treatedby an m-dimensional approximation of the determinant with an actual convergencerate of O(m−2) instead of O(m−1) as for the projection methods. Moreover, thesemethods are much simpler and straightforwardly implemented.

6. Quadrature Methods. In this section we directly approximate the Fredholmdeterminant (1.2) using the Nyström-type method (1.4) that we have motivated atlength in Section 1. We assume throughout that the kernel K is a continuous functionon [a, b]2. The notation simplifies considerably by introducing the n-dimensionalfunctions Kn defined by

Kn(t1, . . . , tn) = det(K(tp, tq)

)np,q=1 .

Their properties are given in Lemma A.4 of the appendix. We then write the Fred-holm determinant shortly as

d(z) = 1 +∞

∑n=1

zn

n!

∫[a,b]n

Kn(t1, . . . , tn) dt1 · · · dtn.

For a given quadrature formula

Q( f ) =m

∑j=1

wj f (xj) ≈∫ b

af (x) dx

we define the associated Nyström-type approximation of d(z) by the expression

dQ(z) = det(δij + z wiK(xi, xj)

)mi,j=1 . (6.1)

The key to error estimates and a convergence proof is the observation that we canrewrite dQ(z) in a form that closely resembles the Fredholm determinant. Namely, byusing the von Koch form (3.6) of matrix determinants, the multi-linearity of minors,and by introducing the n-dimensional product rule (A.3) associated with Q (see theappendix), we get

dQ(z) = 1 +∞

∑n=1

zn

n!

m

∑j1,...,jn=1

det(

wjp K(xjp , xjq))n

p,q=1

= 1 +∞

∑n=1

zn

n!

m

∑j1,...,jn=1

wj1 · · ·wjn det(

K(xjp , xjq))n

p,q=1

= 1 +∞

∑n=1

zn

n!

m

∑j1,...,jn=1

wj1 · · ·wjn Kn(xj1 , . . . , xjn)

= 1 +∞

∑n=1

zn

n!Qn(Kn).

Thus, alternatively to the motivation given in the introductory Section 1, we couldhave introduced the Nyström-type method by approximating each of the multi-dimensional integrals in the power series defining the Fredholm determinant with a

Page 21: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 21

product quadrature rule. Using this form, we observe that the error is given by

dQ(z)− d(z) =∞

∑n=1

zn

n!

(Qn(Kn)−

∫[a,b]n

Kn(t1, . . . , tn) dt1 · · · dtn

), (6.2)

that is, by the exponentially generating function of the quadrature errors for thefunctions Kn. The following theorem generalizes a result that Hilbert (1904, p. 59)had proven for a specific class of quadrature formulae, namely, the rectangular rule.

Theorem 6.1. If a family Qm of quadrature rules converges for continuous functions,then the corresponding Nyström-type approximation of the Fredholm determinant converges,

dQm(z)→ d(z) (m→ ∞),

uniformly for bounded z.Proof. Let z be bounded by M and choose any ε > 0. We split the series (6.2) at

an index N yet to be chosen, getting

|dQm(z)− d(z)| 6N

∑n=1

Mn

n!

∣∣∣∣Qnm(Kn)−

∫[a,b]n

Kn(t1, . . . , tn) dt1 · · · dtn

∣∣∣∣+

∑n=N+1

Mn

n!

∣∣∣∣Qnm(Kn)−

∫[a,b]n

Kn(t1, . . . , tn) dt1 · · · dtn

∣∣∣∣Let Λ be the stability bound of the convergent family Qm of quadrature rules (seeTheorem A.1 of the appendix) and put Λ1 = max(Λ, b− a). Then, by Lemma A.4,the second part of the splitting can be bounded by

∑n=N+1

Mn

n!

∣∣∣∣Qnm(Kn)−

∫[a,b]n

Kn(t1, . . . , tn) dt1 · · · dtn

∣∣∣∣6

∑n=N+1

Mn

n!

(|Qn

m(Kn)|+ |∫[a,b]n

Kn(t1, . . . , tn) dt1 · · · dtn|)

6∞

∑n=N+1

Mn

n!(Λn + (b− a)n) ‖Kn‖L∞ 6 2

∑n=N+1

nn/2

n!(MΛ1‖K‖L∞)n.

The last series converges by Lemma A.5 and the bound can, therefore, be pushedbelow ε/2 by choosing N large enough. After fixing such an N, we can certainlyalso push the first part of the splitting, that is,

N

∑n=1

Mn

n!

∣∣∣∣Qnm(Kn)−

∫[a,b]n

Kn(t1, . . . , tn) dt1 · · · dtn

∣∣∣∣ ,

below ε/2, now for m large enough, say m > m0, using the convergence of theproduct rules Qn

m induced by Qm (see Theorem A.3). In summary we get

|dQm(z)− d(z)| 6 ε

for all |z| 6 M and m > m0, which proves the asserted convergence of the Nyström-type method.

Page 22: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

22 F. BORNEMANN

If the kernel K enjoys, additionally, some smoothness, we can prove error esti-mates that exhibit, essentially, the same rates of convergence as for the quadratureof the sections x 7→ K(x, y) and y 7→ K(x, y).

Theorem 6.2. If K ∈ Ck−1,1([a, b]2), then for each quadrature rule Q of order ν > kwith positive weights there holds the error estimate

|dQ(z)− d(z)| 6 ck 2k(b− a)kν−k Φ(|z|(b− a)‖K‖k) ,

where ck is the constant (depending only on k) from Theorem A.2, and ‖K‖k and Φ are thenorm and function defined in (A.6) and (A.10), respectively.

If K is bounded analytic on Eρ × Eρ (with the ellipse Eρ defined in Theorem A.2), thenfor each quadrature rule Q of order ν with positive weights there holds the error estimate

|dQ(z)− d(z)| 6 4 ρ−ν

1− ρ−1 Φ(|z|(b− a)‖K‖L∞(Eρ×Eρ)

).

Proof. By Theorem A.3 and Lemma A.4 we can estimate the error (6.2) in bothcases in the form

|dQ(z)− d(z)| 6 α∞

∑n=1

n(n+2)/2

n!(|z|β)n = α Φ(|z|β) ;

with the particular values α = ck 2k(b− a)kν−k and β = (b− a) ‖K‖k in the first case,and α = 4 ρ−ν/(1− ρ−1) and β = (b− a) ‖K‖L∞(Eρ×Eρ) in the second case. This provesboth assertions.

An example with an analytic kernel, enjoying the excellent convergence rates ofthe second part this theorem, can be found in Section 7.

Note that Theorem 6.2 is based on a general result (Theorem A.2) about quadra-ture errors that stems from the convergence rates of polynomial best approximation.There are cases (typically of low regularity), however, for which certain quadratureformulae enjoy convergence rates that are actually better than best approximation.The Nyström-type method inherits this behavior; one would just have to repeat theproof of Theorem 6.2 then. We refrain from stating a general theorem, since thiswould involve bounds on the highest derivatives involving weights17 that take intoaccount the boundary of the interval [a, b]. Instead, we content ourselves with thedetailed discussion of a particular example.

An example: Poisson’s equation. We revisit the example of Section 5, that is the in-tegral operator (5.7) belonging to the Green’s kernel K (defined in (5.8)) of Poisson’sequation. Recall from (5.9) that

d(−1) = det(I − A) = sin(1).

The kernel K is just Lipschitz continuous, that is, K ∈ C0,1([0, 1]2). If we apply theNyström-type method with the m-point Gauss–Legendre (order ν = 2m) or theCurtis–Clenshaw (order ν = m) formulae as the underlying quadrature rule Qm,Theorem 6.2 proves an error bound of the form

dQm(−1)− d(−1) = O(m−1),

17For the interval [−1, 1] this weight would be (1− x2)1/2, see Davis and Rabinowitz (1984, §4.8.1).

Page 23: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 23

101

102

103

10−8

10−7

10−6

10−5

10−4

10−3

10−2

m

|dQ

m

(−1)

− d

(−1)

|

Fig. 2. Convergence of the Nyström-type method for approximating the Fredholm determinant of the integraloperator induced by Green’s kernel (5.8) of Poisson’s equation; the underlying quadrature rules Qm are the m-pointGauss–Legendre (dots) and Clenshaw–Curtis (circles) rules. Note that, in accordance with Trefethen (2008), bothbehave essentially the same. The solid line shows the function 1/25m2, just to indicate the rate of convergence. Forcomparison we have included the results of the Ritz–Galerkin method (stars) from Figure 1.

which superficially indicates the same convergence rate as for the m-dimensionalGalerkin methods of Section 5. However, the actual numerical computation shownin Figure 2 exhibits the far better convergence rate of O(m−2). This deviation can beunderstood in detail as follows:

On the one hand, by inverse theorems of approximation theory (DeVore andLorentz 1993, p. 220), valid for proper subintervals of [a, b], the polynomial best ap-proximation (of degree m) of sections of the Green’s kernel K cannot give a betterrate than O(m−1); since otherwise those sections could not show jumps in the firstderivative. Given the line of arguments leading from polynomial best approxima-tion to Theorem 6.2, the error estimate of O(m−1) was therefore the best that couldbe established this way.

On the other hand, the sections of the Green’s kernel look like piecewise linearhat functions. Therefore, the coefficients am of their Chebyshev expansions decayas O(m−2) (Davis and Rabinowitz 1984, Eq. (4.8.1.3)). Given this decay rate, one canthen prove—see, for Gauss–Legendre, Davis and Rabinowitz (1984, Eq. (4.8.1.7)) and,for Clenshaw–Curtis, Riess and Johnson (1972, Eq. (9))—that the quadrature error isof rate O(m−2), too. Now, one can lift this estimate to the Nyström-like methodessentially as in Theorem 6.2; thus proving in fact that

dQm(−1)− d(−1) = O(m−2),

as numerically observed.

Remark. This “superconvergence” property of certain quadrature rules, as op-posed to best approximation, for kernels with jumps in a higher derivative is there-fore also the deeper reason that the Nyström-type method then outperforms theprojection methods of Section 5 (see Figure 2): Best approximation, by direct (Jack-son) and inverse (Bernstein) theorems of approximation theory, is strongly tied withthe regularity of K. And this, in turn, is tied to the decay of the singular values ofthe induced integral operator A, which governs the convergence rates of projectionmethods.

Page 24: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

24 F. BORNEMANN

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 510

−15

10−10

10−5

100

s

E2(0

;s)

Fig. 3. The probability E2(0; s) that an interval of length s does not contain, in the bulk scaling limit oflevel spacing 1, an eigenvalue of the Gaussian unitary ensemble (GUE). The result shown was calculated with theNyström-like method based on Gauss–Legendre with m = 30; and cross-checked against the asymptotic expansionlog E2(0; s) = −π2s2/8− log(s)/4 + log(2)/3− log(π)/4 + 3ζ ′(−1) + O(s−1) for s→ ∞ (Deift et al. 1997).

A note on implementation. If the quadrature weights are positive (which in viewof Theorem A.1 is anyway the better choice), as is the case for Gauss–Legendre andClenshaw–Curtis, we recommend to implement the Nyström-type method (6.1) inthe equivalent, symmetric form

dQ(z) = det(I + zAQ), AQ =(

w1/2i K(xi, xj)w1/2

j

)m

i,j=1. (6.3)

(Accordingly short Matlab and Mathematica code is given in the introductory Sec-tion 1.) The reason is that the m× m-matrix AQ inherits some important structuralproperties from the integral operator A:

• If A is selfadjoint, then AQ is Hermitian (see Footnote 10).

• If A is positive semidefinite, then, by (2.9), AQ is positive semidefinite, too.

This way, for instance, the computational cost for calculating the finite-dimensionaldeterminant is cut to half, if by structural inheritance I + zAQ is Hermitian positivedefinite; the Cholesky decomposition can then be employed instead of Gaussianelimination with partial pivoting.

7. Application to Some Entire Kernels of Random Matrix Theory. In this sec-tion we study two important examples, stemming from random matrix theory, withentire kernels. By Theorem 6.2, the Nyström-type method based on Gauss–Legendreor Curtis–Clenshaw quadrature has to exhibit exponential convergence.

7.1. The sine kernel. The probability E2(0; s) (shown in Figure 3) that an in-terval of length s does not contain, in the bulk scaling limit of level spacing 1, aneigenvalue of the Gaussian unitary ensemble (GUE) is given (Mehta 2004, Sect. 6.3)by the Fredholm determinant

E2(0; s) = det (I − As)

of the integral operator As on L2(0, s) that is induced by the sine kernel K:

Asu(x) =∫ s

0K(x, y)u(y) dy, K(x, y) =

sin(π(x− y))π(x− y)

.

Page 25: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 25

Note that K(x, y) is Hermitian and entire on C×C; thus As is a selfadjoint operatorof trace class on L2(0, s). (This is already much more than we would need to knowfor successfully applying and understanding the Nyström-type method. However,to facilitate a comparison with the Ritz–Galerkin method, we analyze the operatorAs in some more detail.) The factorization

K(x, y) =1

∫ π

−πei(x−y)ξ dξ =

12π

∫ π

−πeixξe−iyξ dξ (7.1)

of the kernel implies that As is positive definite with maximal eigenvalue λ1(As) < 1;since, for 0 6= u ∈ L2(0, s), we obtain

0 < 〈u, Asu〉 =∫ π

−π

∣∣∣∣ 1√2π

∫ s

0e−ixξ u(x) dx

∣∣∣∣2 dξ

=∫ π

−π|u(ξ)|2 dξ <

∫ ∞

−∞|u(ξ)|2 dξ = ‖u‖2

L2 = ‖u‖2L2 .

Here, in stating that the inequalities are strict, we have used the fact that the Fouriertransform u of the function u ∈ L2(0, s), which has compact support, is an entirefunction. Therefore, the perturbation bound of Lemma 4.1 applies and we obtain,for Ritz–Galerkin as for any Galerkin method, like in the example of Section 5, thebasic error estimate

|det(I − Pm AsPm)− det(I − As)| 6 ‖Pm AsPm − As‖J1 .

Now, Theorems 5.2 and 6.2 predict a rapid, exponentially fast convergence of theRitz–Galerkin and the Nyström-type methods: In fact, an m-dimensional approxima-tion will give an error that decays like O(e−cm), for any fixed c > 0 since, for entirekernels, the parameter ρ > 1 can be chosen arbitrarily in these theorems.

Details of the implementation of the Ritz–Galerkin method. There is certainly no gen-eral recipe on how to actually construct the Ritz–Galerkin method for a specificexample, since one would have to know, more or less exactly, the eigenvalues of A.In the case of the sine kernel, however, Gaudin (1961) had succeeded in doing so.(See also Katz and Sarnak (1999, p. 411) and Mehta (2004, pp. 124–126).) He hadobserved that the integral operator At on L2(−1, 1), defined by

Atu(x) =∫ 1

−1eiπtxyu(y) dy

(which is, by (7.1), basically a rescaled “square-root” of A2t), is commuting with theselfadjoint, second-order differential operator

Lu(x) =d

dx

((x2 − 1)u′(x)

)+ π2t2x2u(x)

with boundary conditions

(1− x2)u(x)|x=±1 = (1− x2)u′(x)|x=±1 = 0.

Thus, both operators share the same set of eigenfunctions un, namely the radialprolate spheroidal wave functions (using the notation of Mathematica 6.0)

un(x) = S(1)n,0(πt, x) (n = 0, 1, 2 . . .).

Page 26: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

26 F. BORNEMANN

5 10 15 2010

−20

10−15

10−10

10−5

100

dimension m

appr

oxim

atio

n er

ror

for

E2(0

;1)

5 10 15 20 2510

−20

10−15

10−10

10−5

100

dimension m

appr

oxim

atio

n er

ror

for

E2(0

;2)

Fig. 4. Convergence of various m-dimensional approximations of the Fredholm determinants E2(0; 1) (left)and E2(0; 2) (right): the Nyström-type quadrature methods based on Gauss–Legendre (dots) and Curtis–Clenshaw(circles), as well as Gaudin’s (1961) Ritz–Galerkin method based on spheroidal wave functions (stars). The dashedline shows the amount, according to (4.5), of roundoff error due to the numerical evaluation of the finite-dimensionaldeterminants; all calculations were done in IEEE double arithmetic with machine precision ε = 2−53.

These special functions are even for n even, and odd for n odd. By plugging theminto the integral operator At Gaudin had obtained, after evaluating at x = 0, theeigenvalues

λ2k(At) =1

u2k(0)

∫ 1

−1u2k(y) dy, λ2k+1(At) =

iπtu′2k+1(0)

∫ 1

−1u2k+1(y)y dy.

Finally, we have (starting with the index n = 0 here)

λn(As) =s4|λn(As/2)|2 (n = 0, 1, 2, . . .).

Hence, the m-dimensional Ritz–Galerkin approximation of det(I − As) is given by

det(I − Pm AsPm) =m−1

∏n=0

(1− λn(As)).

While Gaudin himself had to rely on tables of the spheroidal wave functions (Strattonet al. 1956), we can use the fairly recent implementation of these special functionsby Falloon, Abbott and Wang (2003), which now comes with Mathematica 6.0. Thisway, we get the following implementation:

u@t_, n_, x_D := SpheroidalS1@n, 0, π t, xD

μ@t_, n_?EvenQD :=NIntegrate@u@t, n, yD, 8y, −1, 1<D

u@t, n, 0D

μ@t_, n_?OddQD :=π t NIntegrate@y u@t, n, yD, 8y, −1, 1<D

uH0,0,1L@t, n, 0D

λ@s_, n_D := λ@s, nD =s

4AbsBμB

s

2, nFF

2

DetRitzGalerkin@s_, m_D := ‰n=0

m−1

H1 − λ@s, nDL

DetRitzGalerkin@1, 9D

0.1702174213791839

DetRitzGalerkin@2, 11D

0.00349732514916866

Given all this, one can understand that Jimbo et al.’s (1980) beautiful discovery ofexpressing E2(0; s) by formula (1.6) in terms of the fifth Painlevé transcendent wasgenerally considered to be a major break-through even for its numerical evaluation

Page 27: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 27

−8 −7 −6 −5 −4 −3 −2 −1 0 1 210

−20

10−15

10−10

10−5

100

s

F2(s

)

Fig. 5. The Tracy–Widom distribution F2(s); that is, in the edge scaling limit, the probability of the maximaleigenvalue of the Gaussian unitary ensemble (GUE) being not larger than s. The result shown was calculatedwith the Nyström-like method based on Gauss–Legendre with m = 50; and cross-checked against the asymptoticexpansion log F2(−s) = −s3/12− log(s)/8 + log(2)/24 + ζ ′(−1) + O(s−3/2) for s→ ∞ (Deift et al. 2008).

(Tracy and Widom 2000, Footnote 10). However, note how much less knowledgesuffices for the application of the far more general Nyström-type method: continuityof K makes it applicable, and K being entire guarantees rapid, exponentially fastconvergence. That is all.

An actual numerical experiment. Figure 4 shows the convergence (in IEEE machinearithmetic) of an actual calculation of the numerical values E2(0; 1) and E2(0; 2).We observe that the Nyström-type method based on Gauss–Legendre has an expo-nentially fast convergence rate comparable to the Ritz–Galerkin method. Clenshaw–Curtis needs a dimension m that is about twice as large as for Gauss–Legendre toachieve the same accuracy. This matches the fact that Clenshaw–Curtis has the or-der ν = m, which is half the order ν = 2m of Gauss–Legendre, and shows thatthe bounds of Theorem 6.2 are rather sharp with respect to ν (there is no “kink”phenomenon here, cf. Trefethen (2008, p. 84)). The dashed line shows the amount,as estimated in (4.5), of roundoff error that stems from the numerical evaluation ofthe finite dimensional m× m-determinant itself. Note that this bound is essentiallythe same for all the three methods and can easily be calculated in course of thenumerical evaluation. We observe that this bound is explaining the actual onset ofnumerical “noise” in all the three methods reasonably well.

Remark. Note that the Nyström-type method outperforms the Ritz–Galerkin me-thod by far. First, the Nyström-type method is general, simple, and straightforwardlyimplemented (see the code given in Section 1); in contrast, the Ritz–Galerkin dependson some detailed knowledge about the eigenvalues and requires numerical accessto the spheroidal wave functions. Second, there is no substantial gain, as comparedto the Gauss–Legendre based method, in the convergence rate from knowing theeigenvalues exactly. Third, and most important, the computing time for the Ritz–Galerkin runs well into several minutes, whereas both versions of the Nyström-typemethod require just a few milliseconds.

7.2. The Airy kernel. The Tracy–Widom distribution F2(s) (shown in Figure 5),that is, in the edge scaling limit, the probability of the maximal eigenvalue of theGaussian unitary ensemble (GUE) being not larger than s, is given (Mehta 2004,

Page 28: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

28 F. BORNEMANN

§24.2) by the Fredholm determinant

F2(s) = det(I − As) (7.2)

of the integral operator As on L2(s, ∞) that is induced by the Airy kernel K:

Asu(x) =∫ ∞

sK(x, y)u(y) dy, K(x, y) =

Ai(x)Ai′(y)−Ai(y)Ai′(x)x− y

. (7.3)

Note that K is Hermitian and entire on C× C; thus As is selfadjoint. (Again, thisis already already about all we would need to know for successfully applying andunderstanding the Nyström-type method. However, we would like to show that, asfor the sine kernel, the strong perturbation bound of Lemma 4.1 applies to the Airykernel, too.) There is the factorization (Tracy and Widom 1994, Eq. (4.5))

K(x, y) =∫ ∞

0Ai(x + ξ)Ai(y + ξ) dξ,

which relates the Airy kernel with the Airy transform (Vallée and Soares 2004, §4.2)in a similar way as the sine kernel is related by (7.1) with the Fourier transform. Thisproves, because of the super-exponential decay of Ai(x)→ 0 as x → 0, that As is theproduct of two Hilbert–Schmidt operators on L2(s, ∞) and therefore of trace class.Moreover, As is positive semi-definite with maximal eigenvalue λ1(A) 6 1; sinceby the Parseval–Plancherel equality (Vallée and Soares 2004, Eq. (4.27)) of the Airytransform we obtain, for u ∈ L2(s, ∞),

0 6 〈u, Asu〉 =∫ ∞

0

∣∣∣∣∫ ∞

0Ai(x + ξ)u(x) dx

∣∣∣∣2 dξ

6∫ ∞

−∞

∣∣∣∣∫ ∞

0Ai(x + ξ)u(x) dx

∣∣∣∣2 dξ = ‖u‖2L2 .

More refined analytic arguments, or a carefully controlled numerical approximation,show the strict inequality λ1(A) < 1; the perturbation bound of Lemma 4.1 applies.

Modification of the Nyström-type method for infinite intervals. The quadrature meth-ods of Section 6 are not directly applicable here, since the integral operator As isdefined by an integral over the infinite interval (s, ∞). We have the following threeoptions:

1. Using a Gauss-type quadrature formula on (s, ∞) that is tailor-made for thesuper-exponential decay of the Airy function. Such formulae have recentlybeen constructed by Gautschi (2002).

2. Truncating the integral in (7.3) at some point T > s. That is, before usingthe Nyström-type method with a quadrature formula on the finite interval[s, T] (for which the second part of Theorem 6.2 is then applicable, showingexponential convergence), we approximate the Fredholm determinant (7.2)by

det(I − PT AsPT) = det(

I − As�L2(s,T)

),

where the orthonormal projection PT : L2(s, ∞) → L2(s, T), Pu = u · χ[s,T],denotes the multiplication operator by the characteristic function of [s, T].

Page 29: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 29

0 2 4 6 8 10 12 14 16 18 2010

−60

10−50

10−40

10−30

10−20

10−10

100

truncation point T

PT A

s PT −

As in

Hilb

ert−

Sch

mid

t nor

m

Fig. 6. Values of the expression ‖PT AsPT − As‖J2 , which bounds, by (7.4), the error in F2(s) committed bytruncating the integral in (7.3) at a point T > s.

This way we commit an additional truncation error, which has, by passingthrough the perturbation bound of Lemma 4.1, the computable bound

|det(I − PT AsPT)− det(I − As)| 6 ‖PT AsPT − As‖J1 6

‖PT AsPT − As‖J2 =(∫ ∞

T

∫ ∞

T|K(x, y)|2 dxdy

)1/2. (7.4)

Figure 6 shows this bound as a function of the truncation point T. We ob-serve that, for the purpose of calculating (within IEEE machine arithmetic)F2(s) for s ∈ [−8, 2]—as shown in Figure 5—, a truncation point at T = 16would be more than sufficiently safe.

3. Transforming the infinite intervals to finite ones. By using a monotone andsmooth transformation φs : (0, 1) → (s, ∞), defining the transformed inte-gral operator As on L2(0, 1) by

Asu(ξ) =∫ 1

0Ks(ξ, η)u(η) dη, Ks(ξ, η) =

√φ′s(ξ)φ′s(η) K(φs(ξ), φs(η)),

gives the identity

Fs(s) = det(

I − As�L2(s,∞)

)= det

(I − As�L2(0,1)

).

For the super-exponentially decaying Airy kernel K we suggest the transfor-mation

φs(ξ) = s + 10 tan(πξ/2) (ξ ∈ (0, 1)). (7.5)

Note that though K(ξ, η) is a smooth function on [0, 1] it possesses, as afunction on C×C, essential singularities on the lines ξ = 1 or η = 1. Hence,we can only apply the first part of Theorem 6.2 here, which then shows, forGauss–Legendre and Clenshaw–Curtis, a super-algebraic convergence rate,that is, O(m−k) for arbitrarily high algebraic order k. The actual numericalexperiments reported in Figure 7 show, in fact, even exponential conver-gence.

From the general-purpose point of view, we recommend the third option. It isstraightforward and does not require any specific knowledge, or construction, aswould the first and second option.

Page 30: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

30 F. BORNEMANN

0 5 10 15 20 25 30 35 40 45 5010

−20

10−15

10−10

10−5

100

dimension m

appr

oxim

atio

n er

ror

for

F 2(−2)

0 5 10 15 20 25 30 35 40 45 5010

−20

10−15

10−10

10−5

100

dimension m

appr

oxim

atio

n er

ror

for

F 2(−4)

Fig. 7. Convergence of the m-dimensional Nyström-type approximation—using the transformation (7.5)—ofthe Fredholm determinants F2(−2) (left) and F2(−4) (right), based on Gauss–Legendre (dots) and Curtis–Clenshaw(circles). The dashed line shows the amount, according to (4.5), of roundoff error due to the numerical evaluation ofthe finite-dimensional determinants; all calculations were done in IEEE double arithmetic (ε = 2−53).

Remarks on other numerical methods to evaluate F2(s). As for the sine kernel, there isa selfadjoint second-order ordinary differential operator commuting with As (Tracyand Widom 1994, p. 166). Though this has been used to derive some asymptoticformulas, nothing is known in terms of special functions that would enable us tobase a Ritz–Galerkin method on it. As Mehta (2004, p. 453) puts it: “In the case ofthe Airy kernel . . . the differential equation did not receive much attention and itssolutions are not known.”

Prior to our work of calculating F2(s) directly from its determinantal expression,all the published numerical calculations started with Tracy and Widom’s (1994) re-markable discovery of expressing F2(s) in terms of the second Painlevé transcendent;namely

F2(s) = exp(−∫ ∞

s(z− s)q(z)2 dz

)with q(z) being the Hastings–McLeod (1980) solution of Painlevé II,

q′′(z) = 2q(z)3 + z q(z), q(z) ∼ Ai(z) as z→ ∞. (7.6)

Initial value methods for the numerical integration of (7.6) suffer from severe sta-bility problems (Prähofer and Spohn 2004). Instead, the numerically stable way ofsolving (7.6) goes by considering q(z) as a connecting orbit, the other asymptoticstate being

q(z) ∼√−z2

as z→ −∞,

and using numerical two-point boundary value solvers (Dieng 2005).

8. Extension to Systems of Integral Operators. We now consider an N × Nsystem of integrals operators that is induced by continuous kernels Kij ∈ C(Ii × Ij)(i, j = 1, . . . , N), where the Ii ⊂ R denote some finite intervals. The correspondingsystem of integral equations

ui(x) + zN

∑j=1

∫Ij

Kij(x, y)uj(y) dy = fi(x) (x ∈ Ii, i, j = 1, . . . , N) (8.1)

Page 31: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 31

defines, with u = (u1, . . . , uN) and f = ( f1, . . . , fN), an operator equation

u + zAu = f

on the Hilbert space H = L2(I1)⊕ · · · ⊕ L2(IN).

8.1. The Fredholm determinant for systems. Assuming A to be trace class, letus express det(I + zA) in terms of the system (Kij) of kernels. To this end we showthat the system (8.1) is equivalent to a single integral equation; an idea that, essen-tially, can already be found in the early work of Fredholm (1903, p. 388). To simplifynotation, we assume that the Ik are disjoint (a simple transformation of the system ofintegral equations by a set of translations will arrange for this). We then have18

H =N⊕

k=1

L2(Ik) ∼= L2(I), I = I1 ∪ . . . ∪ In.

by means of the natural isometric isomorphism

(u1, . . . , uN) 7→ u =N

∑k=1

χkuk

where χk denotes the characteristic function of the interval Ik. Given this picture, theoperator A can be viewed being the integral operator on L2(I) that is induced by thekernel

K(x, y) =N

∑i,j=1

χi(x)Kij(x, y)χj(y).

By (3.7) we finally get (cf. Gohberg et al. (2000, Thm 6.1))

det(I + zA) =∞

∑n=0

zn

n!

∫In

det(K(tp, tq)

)np,q=1 dt1 · · · dtn

=∞

∑n=0

zn

n!

∫In

(N

∑i1,...,in=1

χi1(t1) · · · χin(tn)

)︸ ︷︷ ︸

=1

det(K(tp, tq)

)np,q=1 dt1 · · · dtn

=∞

∑n=0

zn

n!

N

∑i1,...,in=1

∫Ii1×···×Iin

det(K(tp, tq)

)np,q=1 dt1 · · · dtn

=∞

∑n=0

zn

n!

N

∑i1,...,in=1

∫Ii1×···×Iin

det(

Kipiq(tp, tq))n

p,q=1dt1 · · · dtn.

18The general case could be dealt with by the topological sum, or coproduct, of the intervals Ik ,

N

äk=1

Ik =N⋃

k=1

Ik × {k}.

One would then use (Johansson 2003) the natural isometric isomorphism

H =N⊕

k=1

L2(Ik) ∼= L2

(N

äk=1

Ik

).

Page 32: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

32 F. BORNEMANN

By eventually transforming back to the originally given, non-disjoint intervals Ik, thelast expression is the general formula that we have sought for: det(I + zA) = d(z)with

d(z) =∞

∑n=0

zn

n!

N

∑i1,...,in=1

∫Ii1×···×Iin

det(

Kipiq(tp, tq))n

p,q=1dt1 · · · dtn. (8.2)

This is a perfectly well defined entire function for any system Kij of continuouskernels, independently of whether A is a trace class operator or not. We call it theFredholm determinant of the system.

The determinant of block matrices. In preparation of our discussion of Nyström-type methods for approximating (8.2) we shortly discuss the determinant of N × N-block matrices

A =

A11 · · · A1N

......

AN1 · · · ANN

∈ CM×M, Aij ∈ Cmi×mj , M = m1 + · · ·+ mN .

Starting with von Koch’s formula (3.6), an argument19 that is similar to the one thathas led us to (8.2) yields

det(I + zA) =∞

∑n=0

zn

n!

N

∑i1,...,in=1

mi1

∑k1=1· · ·

min

∑kn=1

det((Aip ,iq)kp ,kq

)n

p,q=1. (8.3)

8.2. Quadrature methods for systems. Given a quadrature formula for each ofthe intervals Ii, namely

Qi( f ) =mi

∑j=1

wij f (xij) ≈∫

Ii

f (x) dx, (8.4)

we aim at generalizing the Nyström-type method of Section 6. We restrict ourselvesto the case of positive weights, wij > 0, and generalize the method from the singleoperator case as given in (6.3) to the system case in the following form:

dQ(z) = det(I + zAQ), AQ =

A11 · · · A1N

......

AN1 · · · ANN

(8.5)

with the sub-matrices Aij defined by the entries

(Aij)p,q = w1/2ip Kij(xip, xjq)w1/2

jq (p = 1, . . . , mi, q = 1, . . . , mj).

19Alternatively, we can use (3.4) and, recursively, the “binomial” formula (Greub 1967, p. 121)

∧k(V0 ⊕V1) =

k⊕j=0

(∧jV0

)⊗(∧k−j

V1

)of exterior algebra, which is valid for general vector spaces V0 and V1.

Page 33: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 33

This can be as straightforwardly implemented as in the case of a single operator.Now, a convergence theory can be built on a representation of the error dQ(z)− d(z)that is analogous to (6.2). To this end we simplify the notation by introducing thefollowing functions on Ii1 × · · · × Iin ,

Ki1,...,in(t1, . . . , tn) = det(

Kipiq(tp, tq))n

p,q=1,

and by defining, for functions f on Ii1 × · · · × Iin , the product quadrature formula

(n

∏k=1

Qik

)( f ) =

mi1

∑j1=1· · ·

min

∑jn=1

wi1 j1 · · ·win jn f (xi1 j1 , . . . , xin jn)

≈∫

Ii1×···×Iin

f (t1, . . . , tn) dt1 · · · dtn.

Thus, we can rewrite the Fredholm determinant (8.2) in the form

d(z) = 1 +∞

∑n=1

zn

n!

N

∑i1,...,in=1

∫Ii1×···×Iin

Ki1,...,in(t1, . . . , tn) dt1 · · · dtn.

Likewise, by observing the generalized von Koch formula (8.3), we put the defini-tion (8.5) of dQ(z) to the form

dQ(z) = 1 +∞

∑n=1

zn

n!

N

∑i1,...,in=1

(n

∏k=1

Qik

)(Ki1,...,in).

Thus, once again, the Nyström–type method amounts for approximating each mul-tidimensional integral of the power series of the Fredholm determinant by usinga product quadrature rule. Given this representation, Theorem 6.2 can straightfor-wardly be generalized to the system case:

Theorem 8.1. If Kij ∈ Ck−1,1(Ii × Ij), then for each set (8.4) of quadrature formulae ofa common order ν > k with positive weights there holds the error estimate

dQ(z)− d(z) = O(ν−k) (ν→ ∞),

uniformly for bounded z.If the Kij are bounded analytic on Eρ(Ii)× Eρ(Ij) (with the ellipse Eρ(Ii) defined, with

respect to Ii, as in Theorem A.2), then for each set (8.4) of quadrature formulae of a commonorder ν with positive weights there holds the error estimate

dQ(z)− d(z) = O(ρ−ν) (ν→ ∞),

uniformly for bounded z.

8.3. Examples from random matrix theory. Here, we apply the Nyström-typemethod (8.5) to two 2× 2-systems of integral operators that have recently been stud-ied in random matrix theory.

Page 34: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

34 F. BORNEMANN

0 2 4 6 8 10 120

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

t

cov(

A2(t

),A

2(0))

Fig. 8. Values of the two-point correlation function cov(A2(t),A2(0)) of the Airy process A2(t) (solid line).The dashed line shows the first term of the asymptotic expansion cov(A2(t),A2(0)) ∼ t−2 as t→ ∞.

Two-point correlation of the Airy process. The Airy process A2(t) describes, in aproperly rescaled limit of infinite dimension, the maximum eigenvalue of Hermi-tian matrix ensemble whose entries develop according to the Ornstein–Uhlenbeckprocess. This stationary stochastic process was introduced by Prähofer and Spohn(2002) and further studied by Johansson (2003). These authors have shown that thejoint probability function is given by a Fredholm determinant; namely

P(A2(t) 6 s1,A2(0) 6 s2) = det

(I −

(A0 At

A−t A0

)�L2(s1,∞)⊕L2(s2,∞)

)(8.6)

with integral operators At that are induced by the kernel functions

Kt(x, y) =

∫ ∞

0e−ξtAi(x + ξ)Ai(y + ξ) dξ, t > 0,

−∫ 0

−∞e−ξtAi(x + ξ)Ai(y + ξ) dξ, otherwise.

(8.7)

Of particular interest is the two-point correlation function

cov(A2(t),A2(0)) = E(A2(t)A2(0))−E(A2(t))E(A2(0)) (8.8)

=∫

R2s1s2

∂2P(A2(t) 6 s1,A2(0) 6 s2)∂s1∂s2

ds1ds2 − c21,

where c1 denotes the expectation value of the Tracy–Widom distribution (7.2). Wehave calculated this correlation function for 0 6 t 6 100 in steps of 0.1 to an absoluteerror of ±10−10, see Figure 8.20 Here are some details about the numerical procedure:

20A table can be obtained from the author upon request. Sasamoto (2005, Fig. 2) shows a plot (whichdiffers by a scaling factor of two in both the function value and the time t) of the closely related function

g2(t) =√

var(A2(t)−A2(0))/2 =√

var(A2(0))− cov(A2(t),A2(0))

—without, however, commenting on either the numerical procedure used or on the accuracy obtained.

Page 35: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 35

• Infinite intervals of integration, such as in the definition (8.7) of the kernelsor for the domain of the integral operators (8.6) themselves, are handled bya transformation to the finite interval [0, 1] as in Section 7.2.

• The kernels (8.7) are evaluated, after transformation, by a Gauss–Legendrequadrature.

• The joint probability distribution (8.6) is then evaluated, after transforma-tion, by the Nyström-type method of this section, based on Gauss–Legendrequadrature.

• To avoid numerical differentiation, the expectation values defining the two-point correlation (8.8) are evaluated by truncation of the integrals, partialintegration, and using a Gauss–Legendre quadrature once more.

Because of analyticity, the convergence is always exponential. With parameters care-fully (i.e., adaptively) adjusted to deliver an absolute error of ±10−10, the evaluationof the two-point correlation takes, for a single time t and using a 2 GHz PC, about 20

minutes on average. The results were cross-checked, for small t, with the asymptoticexpansion (Prähofer and Spohn 2002, Hägg 2007)

cov(A2(t),A2(0)) = var(A2(0))− 12 var(A2(t)−A2(0))

= var(A2(0)) − t + O(t2) (t→ 0),

and, for large t, with the asymptotic expansion21 (Widom 2004, Adler and vanMoerbeke 2005)

cov(A2(t),A2(0)) = t−2 + ct−4 + O(t−6) (t→ ∞),

where the constant c = −3.542 · · · can explicitly be expressed in terms of theHastings–McLeod solution (7.6) of Painlevé II.

Two-point correlation of the Airy1 process. Sasamoto (2005) and Borodin et al. (2007)have introduced the Airy1 process A1(t) for which, once again, the joint probabilitydistribution can be given in terms of a Fredholm determinant; namely

P(A1(t) 6 s1,A1(0) 6 s2) = det

(I −

(A0 At

A−t A0

)�L2(s1,∞)⊕L2(s2,∞)

)with integral operators At that are now induced by the kernel functions

Kt(x, y) =

Ai(x + y + t2)et(x+y)+2t3/3 − exp(−(x− y)2/(4t))√

4πt, t > 0,

Ai(x + y + t2)et(x+y)+2t3/3, otherwise.

21Adler and van Moerbeke (2005) have derived this asymptotic expansion from the masterfully ob-tained result that G(t, x, y) = log P(A2(t) 6 x,A2(0) 6 y) satisfies the following nonlinear 3rd orderPDE with certain (asymptotic) boundary conditions:

t∂

∂t

(∂2

∂x2 −∂2

∂y2

)G =

∂3G∂x2∂y

(2

∂2G∂y2 +

∂2G∂x∂y

− ∂2G∂x2 + x− y− t2

)

− ∂3G∂y2∂x

(2

∂2G∂x2 +

∂2G∂x∂y

− ∂2G∂y2 − x + y− t2

)+(

∂3G∂x3

∂y− ∂3G

∂y3∂

∂x

)(∂

∂x+

∂y

)G.

The reader should contemplate a numerical calculation of the two-point correlation based on this PDE,rather than directly treating the Fredholm determinant as suggested by us.

Page 36: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

36 F. BORNEMANN

0 0.2 0.4 0.6 0.8 1 1.2 1.40

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

t

cov(

A1(t

),A

1(0))

Fig. 9. Values of the two-point correlation function cov(A1(t),A1(0)) of the Airy1 process A1(t).

By basically employing the same numerical procedure as for the Airy process, wehave succeeded in calculating the two-point correlation function cov(A1(t),A1(0))for 0 6 t 6 2.5 in steps of 0.025 to an absolute error of ±10−10, see Figure 9.22 Fora single time t the evaluation takes about 5 minutes on average (using a 2 GHz PC).This numerical result has been used by Bornemann, Ferrari and Prähofer (2008) asa strong evidence that the Airy1 process is, unlike previously conjectured, not thelimit of the largest eigenvalue in GOE matrix diffusion.

A. Appendices.

A.1. Quadrature Rules. For the ease of reference, we collect in this appendixsome classical facts about quadrature rules in one and more dimensions.

Quadrature rules in one dimension. We consider quadrature rules of the form

Q( f ) =m

∑j=1

wj f (xj) (A.1)

which are meant to approximate∫ b

a f (x) dx for continuous functions f on some finiteinterval [a, b] ⊂ R. We define the norm of a quadrature rule by

‖Q‖ =m

∑j=1|wj|

Convergence of a sequence of quadrature rules is characterized by the followingtheorem of Pólya (Davis and Rabinowitz 1984, p. 130).

Theorem A.1. A sequence Qn of quadrature rules converges for continuous functions,

limn→∞

Qn( f ) =∫ b

af (x) dx ( f ∈ C[a, b]),

22A table can be obtained from the author upon request. Sasamoto (2005, Fig. 2) shows a plot (whichdiffers by a scaling factor of two in both the function value and the time t) of the closely related function

g1(t) =√

var(A1(t)−A1(0))/2 =√

var(A1(0))− cov(A1(t),A1(0))

—without, however, commenting on either the numerical procedure used or on the accuracy obtained.

Page 37: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 37

if and only if the sequence ‖Qn‖ of norms is bounded by some stability constant Λ and if

limn→∞

Qn(xk) =∫ b

axk dx (k = 0, 1, 2, . . .). (A.2)

If the weights are all positive, then (A.2) already implies the boundedness of ‖Qn‖ = Qn(1).A quadrature rule Q is of order ν > 1, if it is exact for all polynomials of degree

at most ν− 1. Using results from the theory polynomial best approximation one canprove quite strong error estimates (Davis and Rabinowitz 1984, §4.8).

Theorem A.2. If f ∈ Ck−1,1[a, b], then for each quadrature rule Q of order ν > k withpositive weights there holds the error estimate∣∣∣∣Q( f )−

∫ b

af (x) dx

∣∣∣∣ 6 ck (b− a)k+1ν−k‖ f (k)‖L∞(a,b) ,

with a constant23 ck depending only on k.If f is bounded analytic in the ellipse Eρ with foci at a, b and semiaxes of lengths s > σ

such that

ρ =√

s + σ

s− σ,

then for each quadrature rule Q of order ν with positive weights there holds the error estimate∣∣∣∣Q( f )−∫ b

af (x) dx

∣∣∣∣ 6 4(b− a)ρ−ν

1− ρ−1 ‖ f ‖L∞(Eρ).

Quadrature rules in two and more dimensions. For the n-dimensional integral∫[a,b]n

f (t1, . . . , tn) dt1 · · · dtn

we consider the product quadrature rule Qn that is induced by an one dimensionalquadrature rule Q of the form (A.1), namely

Qn( f ) =m

∑j1,...,jn=1

wj1 · · ·wjn f (xj1 , . . . , xjn). (A.3)

We introduce some further notation for two classes of functions f . First, for f ∈Ck−1,1([a, b]n), we define the seminorm

| f |k =n

∑i=1‖∂k

i f ‖L∞((a,b)n). (A.4)

Second, if f ∈ C([a, b]n) is sectional analytic—that is, analytic with respect to eachvariable ti while the other variables are fixed in [a, b]—in the ellipse Eρ (defined inTheorem A.2), and if f is uniformly bounded there, we call f to be of class Cρ withnorm

‖ f ‖Cρ =n

∑i=1

max(t1,...,ti−1,ti+1,...,tn)∈[a,b]n−1

‖ f (t1, . . . , ti−1, · , ti+1, . . . , tn)‖L∞(Eρ). (A.5)

23Taking Jackson’s inequality as given in Cheney (1998, p. 147), ck = 2(πe/4)k/√

2πk will do the job.

Page 38: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

38 F. BORNEMANN

By a straightforward reduction argument (Davis and Rabinowitz 1984, p. 361) tothe quadrature errors of the one-dimensional coordinate sections of f , Theorems A.1and A.2 can now be generalized to n dimensions.

Theorem A.3. If a sequence of quadrature rules converges for continuous functions,then the same holds for the induced n-dimensional product rules.

If f ∈ Ck−1,1([a, b]n), then for each one-dimensional quadrature rule Q of order ν > kwith positive weights there holds the error estimate∣∣∣∣Qn( f )−

∫[a,b]n

f (t1, . . . , tn) dt1 · · · dtn

∣∣∣∣ 6 ck (b− a)n+k ν−k| f |k ,

with the same constant ck depending only on k as in Theorem A.2.If f ∈ C([a, b]n) is of class Cρ, then for each one-dimensional quadrature rule Q of order

ν with positive weights there holds the error estimate∣∣∣∣Qn( f )−∫[a,b]n

f (t1, . . . , tn) dt1 · · · dtn

∣∣∣∣ 6 4(b− a)nρ−ν

1− ρ−1 ‖ f ‖Cρ .

Notes on Gauss–Legendre and Curtis–Clenshaw quadrature. Arguably, the most in-teresting families of quadrature rules, with positive weights, are the Clenshaw–Curtisand Gauss-Legendre rules. With m points, the first is of order ν = m, the second of or-der ν = 2m. Thus, Theorems A.1 and A.2 apply. The cost of computing the weightsand points of Clenshaw–Curtis is O(m log m) using FFT, that of Gauss–Legendreis O(m2) using the Golub–Welsh algorithm; for details see (Waldvogel 2006) and(Trefethen 2008). The latter paper studies in depth the reasons why the Clenshaw–Curtis rule, despite having only half the order, performs essentially as well as Gauss–Legendre for most integrands. To facilitate reproducibility we offer the Matlab code(which is just a minor variation of the code given in the papers mentioned above)that has been used in our numerical experiments:

function [w,c] = ClenshawCurtis(a,b,m)

m = m-1;

c = cos((0:m)*pi/m);

M = [1:2:m-1]'; l = length(M); n = m-l;

v0 = [2./M./(M-2); 1/M(end); zeros(n,1)];

v2 = -v0(1:end-1)-v0(end:-1:2);

g0 = -ones(m,1); g0(1+l)=g0(1+l)+m; g0(1+n)=g0(1+n)+m;

g = g0/(m^2+mod(m,2));

w = ifft(v2+g); w(m+1) = w(1);

c = ((1-c)/2*a+(1+c)/2*b)';

w = ((b-a)*w/2)';

for Clenshaw–Curtis; and

function [w,c] = GaussLegendre(a,b,m)

k = 1:m-1; beta = k./sqrt((2*k-1).*(2*k+1));

T = diag(beta,-1) + diag(beta,1);

[V,L] = eig(T);

c = (diag(L)+1)/2; c = (1-c)*a+c*b;

w = (b-a)*V(1,:).^2;

Page 39: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 39

for Gauss–Legendre, respectively. Note, however, that the code for Gauss–Legendreis, unfortunately, suboptimal in requiring O(m3) rather than O(m2) operations, sinceit establishes the full matrix V of eigenvectors of the Jacobi matrix T instead ofdirectly calculating just their first components V(1, :) as in the fully fledged Golub–Welsh algorithm. Even then, there may well be more accurate, and more efficient,alternatives of computing the points and weights of Gauss–Legendre quadrature, seethe discussions in Laurie (2001, §2) and Swarztrauber (2002, §4) and the literaturecited therein.

A.2. Determinantal bounds. In Section 6, for a continuous kernel K ∈ C([a, b]2)of an integral operator, we need some bounds on the derivatives of the inducedn-dimensional function

Kn(t1, . . . tn) = det(K(tp, tq)

)np,q=1 .

To this end, if K ∈ Ck−1,1([a, b]2) we define the norm

‖K‖k = maxi+j6k

‖∂i1∂

j2K‖L∞ . (A.6)

Lemma A.4. If K ∈ C([a, b]2), then Kn ∈ C([a, b]n) with

‖Kn‖L∞ 6 nn/2‖K‖nL∞ . (A.7)

If K ∈ Ck−1,1([a, b]2), then Kn ∈ Ck−1,1([a, b]n) with the seminorm (defined in (A.4))

|Kn|k 6 2kn(n+2)/2‖K‖nk . (A.8)

If K is bounded analytic on Eρ × Eρ (with the ellipse Eρ defined in Theorem A.2), then Kn isof class Cρ (defined in (A.5)) and satisfies

‖Kn‖Cρ 6 n(n+2)/2‖K‖nL∞(Eρ×Eρ). (A.9)

Proof. Using the multilinearity of the determinant we have

∂k

∂tki

∂l

∂slj

∣∣∣∣∣∣∣∣∣∣∣∣∣∣

K(t1, s1) · · · K(t1, sj) · · · K(t1, sn)...

......

K(ti, s1) · · · K(ti, sj) · · · K(ti, sn)...

......

K(tn, s1) · · · K(tn, sj) · · · K(tn, sn)

∣∣∣∣∣∣∣∣∣∣∣∣∣∣

=

∣∣∣∣∣∣∣∣∣∣∣∣∣∣

K(t1, s1) · · · ∂l2K(t1, sj) · · · K(t1, sn)

......

...

∂k1K(ti, s1) · · · ∂k

1∂l2K(ti, sj) · · · ∂k

1K(ti, sn)...

......

K(tn, s1) · · · ∂l2K(tn, sj) · · · K(tn, sn)

∣∣∣∣∣∣∣∣∣∣∣∣∣∣,

Page 40: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

40 F. BORNEMANN

which is, by Hadamard’s inequality (Meyer 2000, p. 469),24 bounded by the expres-sion (see also Lax (2002, p. 262))

nn/2(

maxi+j6k+l

‖∂i1∂

j2K‖L∞

)n.

Now, with

∂kj Kn(t1, . . . , tn) =

k

∑l=0

(kl

)∂k−l

∂tk−lj

∂l

∂slj

∣∣∣∣∣∣∣∣K(t1, s1) · · · K(t1, sn)

......

K(tn, s1) · · · K(tn, sn)

∣∣∣∣∣∣∣∣s1=t1,...,sn=tn

we thus get

‖∂kj Kn‖L∞ 6

k

∑l=0

(kl

)nn/2

(maxi+j6k

‖∂i1∂

j2K‖L∞

)n= 2knn/2

(maxi+j6k

‖∂i1∂

j2K‖L∞

)n.

This proves the asserted bounds (A.7) and (A.8) with k = 0 and k > 1, respectively.The class Cρ bound (A.9) follows analogously to the case k = 0.

A.3. Properties of a certain function used in Theorem 6.2. The power series

Φ(z) =∞

∑n=1

n(n+2)/2

n!zn (A.10)

defines an entire function on C (as the following lemma readily implies).Lemma A.5. Let Ψ be the entire function given by the expression

Ψ(z) = 1 +√

π

2z ez2/4

(1 + erf

( z2

)).

If x > 0, then the series Φ(x) is enclosed by:25√eπ

x Ψ(x√

2e) 6 Φ(x) 6 x Ψ(x√

2e).

Proof. For x > 0 we have

Φ(x) = x∞

∑n=1

nn/2

Γ(n)xn−1.

By Stirling’s formula and monotonicity we get for n > 1√eπ6

nn/2

Γ((n + 1)/2) (√

2e )n−16 1;

in fact, the upper bound is obtained for n = 1 and the lower bound for n→ ∞. Thus,by observing

∑n=1

Γ((n + 1)/2)Γ(n)

zn−1 = 1 +√

π

2z ez2/4

(1 + erf

( z2

))= Ψ(z)

we get the asserted enclosure.

24This inequality, discovered by Hadamard in 1893, was already of fundamental importance to Fred-holm’s original theory (Fredholm 1900, p. 41).

25Note the sharpness of this enclosure:√

e/π = 0.93019 · · · .

Page 41: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 41

Acknowledgements. It is a pleasure to acknowledge that this work has startedwhen I attended the programme on “Highly Oscillatory Problems” at the Isaac New-ton Institute in Cambridge. It was a great opportunity meeting Percy Deift there,who introduced me to numerical problems related to random matrix theory (eventhough he was envisioning a general numerical treatment of Painlevé transcendents,but not of Fredholm determinants). I am grateful for his advice as well as for thecommunication with Herbert Spohn and Michael Prähofer who directed my searchfor an “open” numerical problem to the two-point correlation functions of the Airyand Airy1 processes. I thank Patrik Ferrari who pointed me to the (2005) paper ofSasamoto. Given the discovery that Theorem 6.1, which was pivotal to my study, isessentially a long forgotten (see my discussion on p. 4) result of Hilbert’s 1904 work,I experienced much reconciliation—please allow me this very personal statement—from reading the poem “East Coker” (1940), in which T. S. Eliot, that “radical tra-ditionalist”, described the nature of the human struggle for progress in life, art, orscience:

. . . And so each ventureIs a new beginning . . .. . . And what there is to conquerBy strength and submission, has already been discoveredOnce or twice, or several times, by men whom one cannot hopeTo emulate—but there is no competition—There is only the fight to recover what has been lostAnd found and lost again and again . . .

REFERENCES

Ablowitz, M. J. and Fokas, A. S.: 2003, Complex variables: introduction and applications, 2nd edn, CambridgeUniversity Press, Cambridge.

Adler, M. and van Moerbeke, P.: 2005, PDEs for the joint distributions of the Dyson, Airy and sineprocesses, Ann. Probab. 33, 1326–1361.

Albeverio, S. and Høegh-Krohn, R.: 1977, Oscillatory integrals and the method of stationary phase ininfinitely many dimensions, with applications to the classical limit of quantum mechanics. I,Invent. Math. 40, 59–106.

Axler, S.: 1995, Down with determinants!, Amer. Math. Monthly 102, 139–154.Axler, S.: 1997, Linear algebra done right, 2nd edn, Springer-Verlag, New York.Baker, C. T. H.: 1977, The numerical treatment of integral equations, Clarendon Press, Oxford.Birkhoff, G. (ed.): 1973, A source book in classical analysis, Harvard University Press, Cambridge.Bornemann, F., Ferrari, P. L. and Prähofer, M.: 2008, The Airy1 process is not the limit of the largest

eigenvalue in GOE matrix diffusion, arXiv:0806.3410.Borodin, A., Ferrari, P. L., Prähofer, M. and Sasamoto, T.: 2007, Fluctuation properties of the TASEP with

periodic initial configuration, J. Stat. Phys. 129, 1055–1080.Carleman, T.: 1918, Über die Fourierkoeffizienten einer stetigen Function, Acta Math. 41, 377–384.Carleman, T.: 1921, Zur Theorie der linearen Integralgleichungen, Math. Zeitschr. 9, 196–217.Cheney, E. W.: 1998, Introduction to approximation theory, reprint of the 2nd (1982) edn, AMS Chelsea

Publishing, Providence.Courant, R. and Hilbert, D.: 1953, Methods of mathematical physics. Vol. I, Interscience Publishers, Inc., New

York.Davis, P. J. and Rabinowitz, P.: 1984, Methods of numerical integration, 2nd edn, Academic Press, Orlando.Deift, P. A.: 1999, Orthogonal polynomials and random matrices: a Riemann-Hilbert approach, American Mathe-

matical Society, Providence.Deift, P. A., Its, A. R. and Zhou, X.: 1997, A Riemann-Hilbert approach to asymptotic problems arising in

the theory of random matrix models, and also in the theory of integrable statistical mechanics,Ann. of Math. 146, 149–235.

Page 42: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

42 F. BORNEMANN

Deift, P., Its, A. and Krasovsky, I.: 2008, Asymptotics of the Airy-kernel determinant, Comm. Math. Phys.278, 643–678.

Delves, L. M. and Mohamed, J. L.: 1985, Computational methods for integral equations, Cambridge UniversityPress, Cambridge.

DeVore, R. A. and Lorentz, G. G.: 1993, Constructive approximation, Springer-Verlag, Berlin.Dieng, M.: 2005, Distribution Functions for Edge Eigenvalues in Orthogonal and Symplectic Ensembles: Painlevé

Representations, PhD thesis, University of Davis. arXiv:math/0506586v2.Dieudonné, J.: 1981, History of functional analysis, North-Holland Publishing Co., Amsterdam.Dunford, N. and Schwartz, J. T.: 1963, Linear operators. Part II: Spectral theory, John Wiley & Sons.Dyson, F. J.: 1976, Fredholm determinants and inverse scattering problems, Comm. Math. Phys. 47, 171–183.Eastham, M.: 1973, The spectral theory of periodic differential equations, Scottish Academic Press, Edinburgh.Falloon, P. E., Abbott, P. C. and Wang, J. B.: 2003, Theory and computation of spheroidal wavefunctions,

J. Phys. A 36, 5477–5495.Fenyo, S. and Stolle, H.-W.: 1982–1984, Theorie und Praxis der linearen Integralgleichungen. Vol. I–IV,

Birkhäuser, Basel.Fredholm, I.: 1900, Sur une nouvelle méthode pour la résolution du problème de Dirichlet, Öfversigt Kongl.

Vetenskaps-Akad. Förhandlingar 57, 39–46.Fredholm, I.: 1903, Sur une classe d’équations fonctionnelles, Acta Math. 27, 365–390.Fredholm, I.: 1909, Les équations intégrales linéaires, C. R. Congrés des Math. tenu à Stockholm 1909.Gaudin, M.: 1961, Sur la loi limite de l’espacement des valeurs propres d’une matrice aléatoire, Nucl. Phys.

25, 447–458.Gautschi, W.: 2002, Computation of Bessel and Airy functions and of related Gaussian quadrature formu-

lae, BIT 42, 110–118.Gohberg, I. C. and Kreın, M. G.: 1969, Introduction to the theory of linear nonselfadjoint operators, American

Mathematical Society, Providence.Gohberg, I., Goldberg, S. and Kaashoek, M. A.: 1990, Classes of linear operators. Vol. I, Birkhäuser Verlag,

Basel.Gohberg, I., Goldberg, S. and Krupnik, N.: 2000, Traces and determinants of linear operators, Birkhäuser

Verlag, Basel.Golub, G. H. and Van Loan, C. F.: 1996, Matrix computations, 3rd edn, Johns Hopkins University Press,

Baltimore.Greub, W. H.: 1967, Multilinear algebra, Springer-Verlag, New York.Grothendieck, A.: 1956, La théorie de Fredholm, Bull. Soc. Math. France 84, 319–384.Hackbusch, W.: 1995, Integral equations: Theory and numerical treatment, Birkhäuser Verlag, Basel.Hadamard, J.: 1893, Résolution d’une question relative aux dérminants, Bull. Sci. Math. 17, 240–246.Hastings, S. P. and McLeod, J. B.: 1980, A boundary value problem associated with the second Painlevé

transcendent and the Korteweg-de Vries equation, Arch. Rational Mech. Anal. 73, 31–51.Hägg, J.: 2007, Local Gaussian fluctuations in the Airy and discrete PNG processes, arXiv:math/0701880.Higham, N. J.: 2002, Accuracy and stability of numerical algorithms, 2nd edn, Society for Industrial and

Applied Mathematics, Philadelphia.Hilbert, D.: 1904, Grundzüge einer allgemeinen Theorie der linearen Integralgleichungen. (Erste Mit-

teilung), Nachr. Ges. Wiss. Göttingen 1904, 49–91.Hilbert, D.: 1912, Grundzüge einer allgemeinen Theorie der linearen Integralgleichungen, Teubner, Leipzig,

Berlin.Hille, E. and Tamarkin, J. D.: 1931, On the characteristic values of linear integral equations, Acta Math.

57, 1–76.Hochstadt, H.: 1973, Integral equations, John Wiley & Sons, New York.Jimbo, M., Miwa, T., Môri, Y. and Sato, M.: 1980, Density matrix of an impenetrable Bose gas and the fifth

Painlevé transcendent, Phys. D 1, 80–158.Johansson, K.: 2000, Shape fluctuations and random matrices, Comm. Math. Phys. 209, 437–476.Johansson, K.: 2003, Discrete polynuclear growth and determinantal processes, Comm. Math. Phys.

242, 277–329.Jost, R. and Pais, A.: 1951, On the scattering of a particle by a static potential, Physical Rev. 82, 840–851.Katz, N. M. and Sarnak, P.: 1999, Random matrices, Frobenius eigenvalues, and monodromy, American Mathe-

matical Society, Providence.Kline, M.: 1972, Mathematical thought from ancient to modern times, Oxford University Press, New York.Knopp, K.: 1964, Theorie and Anwendung der unendlichen Reihen, 5th edn, Springer-Verlag, Berlin.Kress, R.: 1999, Linear integral equations, 2nd edn, Springer-Verlag, New York.Laurie, D. P.: 2001, Computation of Gauss-type quadrature formulas, J. Comput. Appl. Math. 127, 201–217.Lax, P. D.: 2002, Functional analysis, John Wiley & Sons, New York.

Page 43: Numerical Evaluation of Fredholm Determinants THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 3 if either the eigenvalues of the integral operator are, more or less, explicitly known

ON THE NUMERICAL EVALUATION OF FREDHOLM DETERMINANTS 43

Mehta, M. L.: 2004, Random matrices, 3rd edn, Elsevier/Academic Press, Amsterdam.Meyer, C.: 2000, Matrix analysis and applied linear algebra, Society for Industrial and Applied Mathematics,

Philadelphia.Moiseiwitsch, B.: 1977, Recent progress in atomic collisions theory, Rep. Prog. Phys. 40, 843–904.Nyström, E.: 1930, Über die praktische Auflösung von Integralgleichungen mit Anwendungen auf Rand-

wertaufgaben, Acta Math. 54, 185–204.Pietsch, A.: 2007, History of Banach spaces and linear operators, Birkhäuser, Boston.Plemelj, J.: 1904, Zur Theorie der Fredholmschen Funktionalgleichung, Monatsh. f. Math. 15, 93–128.Porter, D. and Stirling, D. S. G.: 1990, Integral equations, Cambridge University Press, Cambridge.Prähofer, M. and Spohn, H.: 2002, Scale invariance of the PNG droplet and the Airy process, J. Statist.

Phys. 108, 1071–1106.Prähofer, M. and Spohn, H.: 2004, Exact scaling functions for one-dimensional stationary KPZ growth, J.

Statist. Phys. 115, 255–279.Prössdorf, S. and Silbermann, B.: 1991, Numerical analysis for integral and related operator equations,

Birkhäuser Verlag, Basel.Reinhardt, W. P. and Szabo, A.: 1970, Fredholm method. I. A numerical procedure for elastic scattering,

Phys. Rev. A 1, 1162–1169.Rezende, J.: 1994, Feynman integrals and Fredholm determinants, J. Math. Phys. 35, 4357–4371.Riess, R. D. and Johnson, L. W.: 1972, Error estimates for Clenshaw-Curtis quadrature, Proc. 18, 345–353.Sasamoto, T.: 2005, Spatial correlations of the 1D KPZ surface on a flat substrate, J. Phys. A 38, L549–L556.Simon, B.: 1977, Notes on infinite determinants of Hilbert space operators, Advances in Math. 24, 244–273.Simon, B.: 2005, Trace ideals and their applications, 2nd edn, American Mathematical Society, Providence.Smithies, F.: 1937, The eigen-values and singular values of integral equations, Proc. London Math. Soc.

43, 255–279.Smithies, F.: 1958, Integral equations, Cambridge University Press, Cambridge.Spohn, H.: 2008, Personal communication.Stewart, G. W.: 1998, Matrix algorithms. Vol. I: Basic decompositions, Society for Industrial and Applied

Mathematics, Philadelphia.Stratton, J. A., Morse, P. M., Chu, L. J., Little, J. D. C. and Corbató, F. J.: 1956, Spheroidal wave functions,

including tables of separation constants and coefficients, John Wiley & Sons, New York.Swarztrauber, P. N.: 2002, On computing the points and weights for Gauss-Legendre quadrature, SIAM J.

Sci. Comput. 24, 945–954.Tracy, C. A. and Widom, H.: 1994, Level-spacing distributions and the Airy kernel, Comm. Math. Phys.

159, 151–174.Tracy, C. A. and Widom, H.: 2000, Universality of the distribution functions of random matrix theory,

Integrable systems: from classical to quantum (Montréal, QC, 1999), Vol. 26 of CRM Proc. LectureNotes, Amer. Math. Soc., Providence, pp. 251–264.

Trefethen, L. N.: 2008, Is Gauss quadrature better than Clenshaw–Curtis?, SIAM Rev. 50, 67–87.Tricomi, F. G.: 1957, Integral equations, Interscience Publishers, Inc., New York.Vallée, O. and Soares, M.: 2004, Airy functions and applications to physics, Imperial College Press, London.von Koch, H.: 1892, Sur les déterminants infinis et les équations différentielles linéaires, Acta Math. 16, 217–

295.Waldvogel, J.: 2006, Fast construction of the Fejér and Clenshaw-Curtis quadrature rules, BIT 46, 195–202.Webster, A. G.: 1927, Partial differential equations of mathematical physics, G. E. Stechert & Co., New York.Whittaker, E. T. and Watson, G. N.: 1927, A course of modern analysis, 4th edn, Cambridge University Press,

Cambridge.Widom, H.: 2004, On asymptotics for the Airy process, J. Statist. Phys. 115, 1129–1134.Wilkinson, D.: 1978, Continuum derivation of the Ising model two-point function, Phys. Rev. D 17, 1629–

1636.


Recommended